Skip to main content Accessibility help
×
Hostname: page-component-76fb5796d-5g6vh Total loading time: 0 Render date: 2024-04-28T12:12:08.968Z Has data issue: false hasContentIssue false

Part II - Marsh Dynamics

Published online by Cambridge University Press:  19 June 2021

Duncan M. FitzGerald
Affiliation:
Boston University
Zoe J. Hughes
Affiliation:
Boston University
Get access

Summary

Image of the first page of this content. For PDF version, please use the ‘Save PDF’ preceeding this image.'
Type
Chapter
Information
Salt Marshes
Function, Dynamics, and Stresses
, pp. 155 - 334
Publisher: Cambridge University Press
Print publication year: 2021

Access options

Get access to the full version of this content by using one of the access options below. (Log in options will check for institutional or personal access. Content may require purchase if you do not have access.)

References

References

Alizad, K., Hagen, S. C., Morris, J. T., Medeiros, S. C., Bilskie, M. V., and Weishampel, J. F. 2016. Coastal wetland response to sea level rise in a fluvial estuarine system. Earth’s Future, 4. doi:10.1002/2016EF000385Google Scholar
Allen, J. R. L. 1990. Saltmarsh growth and stratification: a numerical model with special reference to the Severn Estuary, southwest Britain. Marine Geology, 95: 796.Google Scholar
Barbier, E. B., Hacker, S. D. Kennedy, C., Koch, E. W., Stier, A. C., and Silliman, B. R. 2011. The value of estuarine and coastal ecosystem services. Ecological Monographs, 81: 169193.Google Scholar
Baumfalk, Y. A. 1979. On the pumping activity of Arenicola marina. Netherlands Journal of Sea Research, 13: 422427.Google Scholar
Benner, R., Fogel, M. L., Sprague, E. K., and Hodson, R. E. 1987. Depletion of 13C in lignin and its implications for stable carbon isotope studies. Nature, 329: 708710.Google Scholar
Benner, R., and Hodson, R. E. 1985. Microbial degradation of the leachable and lignocellulosic components of leaves and wood from Rhizophora mangle in a tropical mangrove swamp. Marine Ecology Progress Series, 23: 221230.Google Scholar
Buth, G. J. C., and Voesenek, L. A. C. J. 1987. Decomposition of standing and fallen litter of halophytes in a Dutch salt marsh. In: Huiskes, A. H. L., Blom, C. W. P. M., and Rozema, J., eds., Geobotany 11: Vegetation Between Land and Sea. Dr. W. Junk Publishers, Dordrecht, pp. 146165.CrossRefGoogle Scholar
Cahoon, D., and Guntenspergen, G. 2010. Climate change, sea-level rise, and coastal wetlands. National Wetlands Newsletter, 32: 812.Google Scholar
Cahoon, D. R., White, D. A., and Lynch, J. C. 2011. Sediment infilling and wetland formation dynamics in an active crevasse splay of the Mississippi River delta. Geomorphology, 131: 5768.Google Scholar
Callaway, J. C., Borgnis, E. L., Turner, R. E., and Milan, C. S. 2012. Carbon sequestration and sediment accretion in San Francisco Bay tidal wetlands. Estuaries & Coasts, 35: 11631181.Google Scholar
Callaway, J. C., Nyman, J. A., and DeLaune, R. D. 1996. Sediment accretion in coastal wetlands: a review and a simulation model of processes. Current Topics in Wetland Biogeochemistry, 2: 223.Google Scholar
Callaway, J. C., Schile, L. M., Borgnis, E. L., Busnardo, M., Archbald, G., and Duke, R. 2013. Sediment dynamics and vegetation recruitment in newly restored salt ponds – Final Report for Pond A6 Sediment Study: Submitted to Resources Legacy Fund and South Bay Salt Pond Restoration Project, 29 p., www.southbayrestoration.org/documents/technical/Pond%20A6%20FINAL%20report.COMBINED.08.21.2013.pdfGoogle Scholar
Ciutat, A., Widdows, J., and Readman, J. W. 2006. Influence of cockle Cerastoderma edule bioturbation and tidal-current cycles on resuspension of sediment and polycyclic aromatic hydrocarbons. Marine Ecology Progress Series, 328: 5164.Google Scholar
Costanza, R., Perez-Maqueo, O., Martinez, M. L., Sutton, P. Anderson, S., and Mulder, K. 2008. The value of coastal wetlands for hurricane protection. Ambio, 37: 241248.CrossRefGoogle ScholarPubMed
Craft, C., Clough, J. Ehman, J. Joye, S. Park, R., Pennings, S., Guo, H., and Machmuller, M. 2009. Forecasting the effects of accelerated sea-level rise on tidal marsh ecosystem services. Frontiers in Ecology, 7: 7378.Google Scholar
Craft, C. B., Megonigal, J. P., Broome, S. W., Cornell, J., Freese, R., Stevenson, R. J., Zheng, L., and Sacco, J. 2003. The pace of ecosystem development of constructed Spartina alterniflora marshes. Ecological Applications, 13: 14171432.Google Scholar
Craft, C. B., Seneca, E. D., and Broome, S. W. 1993. Vertical accretion in regularly and irregularly flooded microtidal estuarine marshes. Estuarine Coastal and Shelf Science, 37: 371386.Google Scholar
Dangendorf, S., Marcos, M., Wöppelmann, G., Conrad, C. P., Frederikse, T., and Riva, R. 2017. Reassessment of 20th century global mean sea level rise. Proceedings of the National Academy of Sciences of the USA, 114: 59465951.Google Scholar
Darby, F. A., and Turner, R. E. 2008. Below- and aboveground biomass of Spartina alterniflora: Response to nutrient addition in a Louisiana salt marsh. Estuaries and Coasts, 31: 326334.CrossRefGoogle Scholar
Davidson-Arnott, R. G. D., van Proosdij, D., Ollerhead, J., and Schostak, L. E. 2002. Hydrodynamics and sedimentation in salt marshes: examples from a macro-tidal marsh, Bay of Fundy. Geomorphology, 48: 209231CrossRefGoogle Scholar
Davis, J. L., Currin, C. A., O’Brien, C., Raffenburg, C., and Davis, A. 2015. Living shorelines: Coastal resilience with a blue carbon benefit. PLOS ONE, 10(11): e0142595. doi: 10.1371/journal.pone.0142595Google Scholar
de Deckere, E. M. G. T., Tolhurst, T. J., and de Brouwera, J. F. C. 2001. Destabilization of cohesive intertidal sediments by infauna. Estuarine, Coastal and Shelf Science, 53: 665-669.CrossRefGoogle Scholar
Dijkema, K. S., Bossinade, J. H., Bouwsema, P., and de Glopper, R. J. 1990. Salt marshes in the Netherlands Wadden Sea: Rising high-tide levels and accretion enhancement. In: Beukema, J. J., Wolff, W. J, and Brouns, J. J. W. M., eds., Developments in Hydrobiology 57: Expected Effects of Climatic Change on Marine Coastal Ecosystems. Springer, Dordrecht, pp. 173188.Google Scholar
Drake, K., Halifax, H., Adamowicz, S. C., and Craft, C. 2015. Carbon sequestration in tidal salt marshes of the northeast United States. Environmental Management, 56: 9981008.CrossRefGoogle ScholarPubMed
Esselink, P., and Zwarts, L. 1989. Seasonal trend in burrow depth and tidal variation in feeding activity of Nereis diversicolor. Marine Ecology Progress Series, 56: 243254.Google Scholar
Federer, C. A., Turcotte, D. E., and Smith, C. T. 1993. The organic fraction–bulk density relationship and the expression of nutrient content in forest soils. Canadian Journal of Forest Research, 23: 10261032.Google Scholar
French, J. R. 1993. Numerical simulation of vertical marsh growth and adjustment to accelerated sea-level rise, North Norfolk, UK. Earth Surface Processes and Landforms, 18: 6381.Google Scholar
He, Y., Widney, S. E., Ruan, M., Herbert, E. R., Li, X., and Craft, C. 2016. Accumulation of soil carbon drives denitrification potential and laboratory-incubated greenhouse gas flux along a chronosequence of salt marsh development. Estuarine, Coastal and Shelf Science, 172: 7280.Google Scholar
Hodson, R. E., Christian, R. R., and Maccubbin, A. E. 1984. Lignocellulose and lignin in the salt marsh grass Spartina alterniflora: initial concentrations and short-term, post-depositional changes in detrital matter. Marine Biology, 81: 17.CrossRefGoogle Scholar
Kirwan, M. L., Guntenspergen, G. R., D’Alpaos, A., Morris, J. T., Mudd, S. M., and Temmerman, S. 2010. Limits on the adaptability of coastal marshes to rising sea level. Geophysical Research Letters, 37: L23401. doi: 10.1029/2010GL045489.Google Scholar
Kirwan, M. L., and Murray, A. B. 2007. A coupled geomorphic and ecological model of tidal marsh evolution. Proceedings of the National Academy of Sciences of the USA, 104: 61186122,Google Scholar
Krone, R. B. 1987. A method for simulating historic marsh elevations. In: Krause, N. C., ed., Coastal Sediments ′87: American Society of Civil Engineers, New York, NY, pp. 316323.Google Scholar
Krüger, F. 1964. Messungen der Pumptätigkeit von Arenicola marina L. im Watt. Helgoländer wissenschaftliche Meeresuntersuchungen, 11: 7091.Google Scholar
Leonard, L. A. 1997. Controls on sediment transport and deposition in an incised mainland marsh basin, southeastern North Carolina. Wetlands, 17: 263274.Google Scholar
Lovelock, C. E., Bennion, V., Grinham, A., and Cahoon, D. R. 2011. The role of surface and subsurface processes in keeping pace with sea level rise in intertidal wetlands of Moreton Bay, Queensland, Australia. Ecosystems, 14: 745757.Google Scholar
Marani, M, Da Lio, C., and D’Alpaos, A. 2013. Vegetation engineers marsh morphology through multiple competing stable states. Proceedings of the National Academy of Sciences of the USA, 110: 32593263.Google Scholar
Marion, C., Anthony, E. J., and Trentesaux, A. 2009. Short-term (≤2 yrs) estuarine mudflat and saltmarsh sedimentation: High-resolution data from ultrasonic altimetery, rod surface-elevation table, and filter traps. Estuarine, Coastal and Shelf Science, 83: 475484.Google Scholar
McKee, K. L., and Patrick, W. Jr 1988. The relationship of smooth cordgrass (Spartina alterniflora) to tidal datums: a review. Estuaries, 11: 143151.Google Scholar
Mendelssohn, I. A., and Morris, J. T. 2000. Ecophysiological controls on the growth of Spartina alterniflora. In: Weinstein, N. P., and Kreeger, D. A., eds., Concepts and Controversies in Tidal Marsh Ecology. Kluwer Academic Publishers, Dordrecht, pp. 5980.Google Scholar
Morris, J. T. 2006. Competition among marsh macrophytes by means of geomorphological displacement in the intertidal zone. Estuarine and Coastal Shelf Science, 69: 395402.Google Scholar
Morris, J. T. 2007. Ecological engineering in intertidal saltmarshes. Hydrobiologia, 577: 161168.Google Scholar
Morris, J. T., Barber, D. C., Callaway, J.C., Chambers, R., Hagen, S. C., Hopkinson, C. S., Johnson, B. J. et al. 2016. Contributions of organic and inorganic matter to sediment volume and accretion in tidal wetlands at steady state, Earth’s Future, 4. doi: 10.1002/2015EF000334.Google Scholar
Morris, J. T., and Bowden, , W. B. 1986. A mechanistic, numerical model of sedimentation, mineralization, and decomposition for marsh sediments. Soil Science Society of America Journal, 50: 96105.Google Scholar
Morris, J. T., Edwards, J., Crooks, S., and Reyes, E. 2012. Assessment of carbon sequestration potential in coastal wetlands. In Lal, R., Lorenz, K., Hüttl, R., Schneider, B. U., and von Braun, J., eds., Recarbonization of the Biosphere: Ecosystem and Global Carbon Cycle. Springer, Dordrecht, pp. 517531.CrossRefGoogle Scholar
Morris, J. T., Sundareshwar, P. V., Nietch, C. T., Kjerfve, B., and Cahoon, D. R. 2002. Responses of coastal wetlands to rising sea level. Ecology, 83: 28692877.CrossRefGoogle Scholar
Morris, J. T., Sundberg, K., and Hopkinson, C. S. 2013. Salt marsh primary production and its responses to relative sea level and nutrients in estuaries at Plum Island, Massachusetts, and North Inlet, South Carolina, USA. Oceanography, 26: 7884.Google Scholar
Mudd, S. M., D’Alpaos, A., and Morris, J. T. 2010. How does vegetation affect sedimentation on tidal marshes? Investigating particle capture and hydrodynamic controls on biologically mediated sedimentation. Journal of Geophysical Research-Earth Surface, 115, F03029, doi: 10.1029/2009JF001566.Google Scholar
Mudd, S. M., Fagherazzi, S., Morris, J. T., and Furbish, D. J. 2004. Flow, sedimentation, and biomass production on a vegetated salt marsh in South Carolina: toward a predictive model of marsh morphologic and ecologic evolution. In: Fagherazzi, S., Marani, A., and Blum, L. K., eds., The Ecogeomorphology of Tidal MarshesAmerican Geophysical Union, Washington, DC, pp. 165187.Google Scholar
Mudd, S. M., Howell, S., and Morris, J. T. 2009. Impact of the dynamic feedback between sedimentation, sea level rise, and biomass production on near surface marsh stratigraphy and carbon accumulation. Estuarine, Coastal and Shelf Science, 82: 377389.Google Scholar
Narayan, S., Beck, M. W., Wilson, P., Thomas, C. J., Guerrero, A., Shepard, C. C., Reguero, B. G., Franco, G., Ingram, J. C., and Trespalacios, D. 2017. The value of coastal wetlands for flood damage reduction in the northeastern USA. Scientific Reports, 7. doi: 10.1038/s41598-017-09269-z.Google Scholar
Neubauer, S. C. 2008. Contributions of mineral and organic components to tidal freshwater marsh accretion. Estuarine, Coastal and Shelf Science, 78: 7888.CrossRefGoogle Scholar
Neubauer, S. C., Anderson, I. C., Constantine, J. A., and Kuehl, S. A. 2002. Sediment deposition and accretion in a mid-Atlantic (U.S.A.) tidal freshwater marsh. Estuarine, Coastal and Shelf Science, 54: 713727.Google Scholar
Nyman, J. A., Delaune, R. D., Walters, R. J., and Patrick, W. H. Jr 1993. Relationship between vegetation and soil formation in a rapidly submerging coastal marsh. Marine Ecology Progress Series, 96: 269279.Google Scholar
Oliver, F.W. 1925. Spartina townsendii: Its mode of establishment, economic uses and taxonomic status. Journal of Ecology, 13: 7491Google Scholar
Pethick, J. 1981. Long-term accretion rates on tidal salt marshes. Journal of Sedimentary Petrology, 51: 571577.Google Scholar
Randerson, P. F. 1979. A simulation model of salt-marsh development and plant ecology.  In: Knights, B. and Phillips, A. J., eds., Estuarine and Coastal Reclamation and Water Storage. Saxon House, Farnborough, pp. 4867.Google Scholar
Ranwell, D. S. 1964. Spartina salt marshes in Southern England. II: Rate and seasonal pattern of sediment accretion. Journal of Ecology, 52: 7994.Google Scholar
Reddy, K. R., Osborne, T. Z., Inglett, K. S., and Corstanje, R. 2006. Influence of water levels on subsidence of organic soils in the upper St. Johns River basin. Final Report SH45812, St. Johns Water Management District. Palatka, FL.Google Scholar
Richard, G. A. 1978. Seasonal and environmental variations in sediment accretion in a Long Island salt marsh. Estuaries, 1: 2935.Google Scholar
Schile, L. M., Callaway, J. C., Morris, J. T., Stralberg, D., Parker, V. T., and Kell, M. 2014. Modeling tidal wetland distribution with sea-level rise: Evaluating the role of vegetation, sediment, and upland habitat in marsh resiliency. PLOS ONE, 9: e88760. doi: 10.1371/journal.pone.0088760Google Scholar
Shepard, C. C., Crain, C. M., and Beck, M. W. 2011. The protective role of coastal marshes: A systematic review and meta-analysis. PLOS ONE, 6: e27374.Google Scholar
Stapf, O. 1907. Mud binding grasses. Bulletin of Miscellaneous Information (Royal Botanic Gardens, Kew), 1907: 190197.Google Scholar
Stewart, V. I., Adams, W. A., and Abdulla, H. H. 1970. Quantitative pedological studies on soils derived from Silurian mudstones. Journal of Soil Science, 21: 248255.Google Scholar
Temmerman, S., Govers, G., Wartel, S., and Meire, P. 2004. Modelling estuarine variations in tidal marsh sedimentation: response to changing sea level and suspended sediment concentrations. Marine Geology, 211: 119.Google Scholar
Thompson, P. R., Hamlington, B. D., Landerer, F. W., and Adhikari, S. 2016. Are long tide gauge records in the wrong place to measure global mean sea level rise? Geophysics Research Letters, 43: 403410,411.Google Scholar
Torres, J. J., Gluck, D. L., and Childress, J. J. 1977. Activity and physiological significance of the pleopods in the respiration of Callianassa californiensis (Dana) (Crustacea: Thalassinidea). Biological Bulletin, 152: 134146.Google Scholar
Turner, R. E., Swenson, E. M., and Milan, C. S. 2000. Organic and inorganic contributions to vertical accretion in salt marsh sediments. In: Weinstein, M. P., and Kreeger, D. A., eds., Concepts and Controversies in Tidal Marsh Ecology, Kluwer Academic Publishers, Dordrecht, pp. 583595.Google Scholar
Vader, W. J. M. 1964. A preliminary investigation into the reactions of the infauna of the tidal flats to tidal fluctuations in water level. Netherlands Journal of Sea Research, 2: 189222.Google Scholar
van Proosdij, D., Ollerhead, J., and Davidson-Arnott, R. G. D. 2006. Seasonal and annual variations in the sediment mass balance of a macro-tidal salt marsh. Marine Geology, 225: 103127.Google Scholar
Volkenborn, N., Polerecky, L., Wethey, D. S., and Woodin, S. A. 2010. Oscillatory porewater bioadvection in marine sediments induced by hydraulic activities of Arenicola marina. Limnology and Oceanography, 55: 12311247.Google Scholar
Webb, A. P., and Eyre, B. D. 2004. Effect of natural populations of burrowing thalassinidean shrimp on sediment irrigation, benthic metabolism, nutrient fluxes and denitrification. Marine Ecology Progress Series, 268: 205220.Google Scholar
Weston, N. B. 2014. Declining sediments and rising seas: an unfortunate convergence for tidal wetlands, Estuaries and Coasts, 37: 123.Google Scholar
Widdows, R. J., Blauw, A., Heip, C. H. R., Herman, P. M. J., Lucas, C. H., Middelburg, J. J., Schmidt, S., Brinsley, M. D., Twisk, F., and Verbeek, H. 2004. Role of physical and biological processes in sediment dynamics of a tidal flat in Westerschelde Estuary, SW the Netherlands. Marine Ecology Progress Series, 274: 4156.Google Scholar
Williams, P. B., and Orr, M. K. 2002. Physical evolution of restored breached levee salt marshes in the San Francisco Bay estuary. Restoration Ecology, 10: 527542Google Scholar
Wilson, J. O., Buchsbaum, R., Valiela, I., and Swain, T. 1986. Decomposition in salt marsh ecosystems: phenolic dynamics during decay of litter of Spartina alterniflora. Marine Ecology Progress Series, 29: 177187.CrossRefGoogle Scholar

References

Addino, M. S., Montemayor, D. I., Escapa, M., Alvarez, M. F., Valiñas, M. S., Lomovasky, B. J., and Iribarne, O. 2015. Effect of Spartina alterniflora Loisel, 1807 on growth of the stout razor clam Tagelus plebeius (Lightfoot, 1786) in a SW Atlantic estuary. Journal of Experimental Marine Biology and Ecology, 463: 135142.Google Scholar
Alber, M., Swenson, E. M., Adamowicz, S. C., and Mendelssohn, I. 2008. Salt marsh dieback: An overview of recent events in the US. Estuarine, Coastal and Shelf Science, 80: 111.Google Scholar
Alberti, J., Escapa, M., Daleo, P., Iribarne, O., Silliman, B., and Bertness, M. 2007. Local and geographic variation in grazing intensity by herbivorous crabs in SW Atlantic salt marshes. Marine Ecology Progress Series, 349: 235243.Google Scholar
Alberti, J., Escapa, M., Iribarne, O., Silliman, B., and Bertness, M. 2008. Crab herbivory regulates plant facilitative and competitive processes in Argentinean marshes. Ecology, 89: 155164.CrossRefGoogle ScholarPubMed
Allen, E., and Curran, H. A., 1974. Biogenic sedimentary structures produced by crabs in lagoon margin and salt marsh environments near Beaufort, North Carolina. Journal of Sedimentary Research, 44: 538548.Google Scholar
Allen, J. R. L. 2000. Morphodynamics of Holocene salt marshes: A review sketch from the Atlantic and Southern North Sea coasts of Europe. Quaternary Science Reviews, 19: 11551231.CrossRefGoogle Scholar
Altieri, A. H., Silliman, B. R., and Bertness, M. D. 2007. Hierarchical organization via a facilitation cascade in intertidal cordgrass bed communities. The American Naturalist, 169: 195206Google Scholar
Anderson, M. E., Smith, J. M., and McKay, S. K. 2011. Wave dissipation by vegetation. USACOE Technical Report AD1003881.Google Scholar
Angelini, C., Griffin, J. N., Van de Koppel, J., Lamers, L. P. M., Smolders, A. J. P., Derksen-Hooijberg, M., van der Heide, T., and Silliman, B. R. 2016. A keystone mutualism underpins resilience of a coastal ecosystem to drought. Nature Communications, doi:10.1038/ncomms12473Google Scholar
Angelini, C., Heide, T., Griffin, J. N., Morton, J. P., Derksen-Hooijberg, M., Lamers, L. P. M., Smolders, A. J. P., and Silliman, B. R. 2015. Foundation species’ overlap enhances biodiversity and multifunctionality from the patch to landscape scale in southeastern United States salt marshes. Proceedings of the Royal Society of London B, 282: 20150421.Google Scholar
Augustin, L. N., Irish, J. L., and Lynett, P. 2009. Laboratory and numerical studies of wave damping by emergent and near-emergent wetland vegetation. Coastal Engineering, 56: 332340.CrossRefGoogle Scholar
Austen, I., Andersen, T. J., and Edelvang, K. 1999. The influence of benthic diatoms and invertebrates on the erodibility of an intertidal mudflat, the Danish Wadden Sea. Estuarine Coastal and Shelf Science, 49: 99111.Google Scholar
Bartholdy, J. 2012. Salt marsh sedimentation. In: Davis, R. A., and Dalrymple, R. W., eds, Principals of Tidal Sedimentology, pp. 151–185.Google Scholar
Bayne, B. L. 2017. Biology of oysters. Developments in Aquaculture and Fisheries Science. Vol. 41: 2844.Google Scholar
Benner, R., Fogel, M. L., and Sprague, E. K. 1991. Diagenesis of belowground biomass of Spartina alterniflora in salt‐marsh sediments. Limnology and Oceanography, 36: 13581374.Google Scholar
Bertness, M. D. 1984a. Habitat and community modification by an introduced herbivorous snail. Ecology, 65: 370381.Google Scholar
Bertness, M. D. 1984b. Ribbed mussels and Spartina alterniflora production in a New England salt marsh. Ecology, 65: 17941807.Google Scholar
Bertness, M. D. 1985. Fiddler crab regulation of Spartina alterniflora production on a New England salt marsh. Ecology, 66: 10421055.CrossRefGoogle Scholar
Bertness, M. D., Gough, L., and Shumway, S. 1992. Salt tolerances and the distribution of fugitive salt marsh plants. Ecology, 73: 18421851.Google Scholar
Bertness, M. D., Holdredge, C., and Altieri, A. H. 2009. Substrate mediates consumer control of salt marsh cordgrass on Cape Cod, New England. Ecology, 90: 21082117.Google Scholar
Bertness, M. D., and Miller, T. 1984. The distribution and dynamics of Uca pugnax (Smith) burrows in a New England salt marsh. Journal of Experimental Marine Biology and Ecology, 83: 211237.Google Scholar
Bilkovic, D. M., Mitchell, M.M., Isdell, R. E., Schliep, M., and Smyth, A. R. 2017. Mutualism between ribbed mussels and cordgrass enhances salt marsh nitrogen removal. Ecosphere, 8: e01795. 10.1002/ecs2.1795Google Scholar
Blum, L. K., and Davey, E. 2013. Below the saltmarsh surface: visualization of plant roots by computer-aided tomography. Oceanography, 26: 8587.Google Scholar
Boorman, L. A, Garbutt, A., and Barratt, D. 1998. The role of vegetation in determining patterns of the accretion of salt marsh sediment. In: Black, K. S., Paterson, D. M., and Cramp, A., eds, Sedimentary Processes in the Intertidal Zone. Geological Society (London), Special Publication No 139, pp. 389399.Google Scholar
Botto, F., and Iribarne, O. 2000. Contrasting effects of two burrowing crabs (Chasmagnathus granulata and Uca uruguayensis) on sediment composition and transport in estuarine environments. Estuarine, Coastal and Shelf Science, 51: 141151.Google Scholar
Botto, F., Iribarne, O., Gutierres, J., Bava, J., Gagliardini, A., and Valiela, I. 2006. Ecological importance of passive deposition of organic matter into burrows of the SW Atlantic crab Chasmagnathus granulatus. Marine Ecology Progress Series, 312: 201210.Google Scholar
Boudreau, B. P., and Imboden, D. M. 1987. Mathematics of tracer mixing in sediments III: The theory of nonlocal mixing within sediments. American Journal of Science, 287: 693719.Google Scholar
Bouma, T. J., Friedrichs, M., Van Wesenbeeck, B. K., Temmerman, S., Graf, G., and Herman, P. M. J. 2009. Density-dependent linkage of scale-dependent feedbacks : A flume study on the intertidal macrophyte Spartina anglica. Oikos, 118: 260268.Google Scholar
Cadee, G. C. 2001. Sediment dynamics by bioturbating organisms. In: Ecological Comparisons of Sedimentary Shores, Springer-Verlag, Berlin, pp. 127147.Google Scholar
Chapman, V. J. 1960. Salt Marshes and Salt Deserts of the World. Leonard Hill: London.Google Scholar
Chmura, G. L., and Kosters, E. C. 1994. Storm deposition and 137Cs accumulation in fine-grained marsh sediments of the Mississippi Delta Plain. Estuarine Coastal and Shelf Science, 39: 3344.Google Scholar
Christiansen, T., Wiberg, P. L., and Milligan, T. G. 2000. Flow and sediment transport on a tidal salt marsh surface. Estuarine, Coastal and Shelf Science, 50: 315331.CrossRefGoogle Scholar
Coverdale, T. C., Altieri, A. H. and Bertness, M. D. 2012. Belowground herbivory increases vulnerability of New England salt marshes to die-off. Ecology, 93: 20852094.CrossRefGoogle ScholarPubMed
Collins, L. M., Collins, J. N., and Leopold, L. B. 1987. Geomorphic processes of an estuarine marsh: preliminary hypotheses. In: Gardiner, V., ed., International Geomorphology, Part 1, Wiley, Chichester, pp. 10491072.Google Scholar
Costanza, R., Pérez-Maqueo, O., Martinez, M. L., Sutton, P., Anderson, S. J., and Mulder, K. 2008. The value of coastal wetlands for hurricane protection. AMBIO: A Journal of the Human Environment, 37: 241248.CrossRefGoogle ScholarPubMed
Crawford, F. 2018. Geomorphology of shell ridges and their effect on the stabilization of Biloxi marshes, east Louisiana. MS thesis, University of New Orleans.Google Scholar
Cuadrado, D. G., Perillo, G. M. E. and Vitale, A. J. 2014. Modern microbial mats in siliciclastic tidal flats: Evolution, structure and the role of hydrodynamics. Marine Geology, 352: 367380Google Scholar
Currin, C. A., Chappell, W. S., and Deaton, A. 2010. Developing alternative shoreline armoring strategies: The living shoreline approach in North Carolina. In: Shipman, H., Dethier, M. N., Gelfenbaum, G., Fresh, K. L., and Dinicola, R. S., eds., Puget Sound Shorelines and the Impacts of Armoring – Proceedings of a State of the Science Workshop, May 2009, Reston, Virginia: U.S. Geological Survey, Scientific Investigations Report 2010-5254, pp. 91102.Google Scholar
Daborn, G. R., Amos, C. L., Brylinsky, M., Christian, H., Drapeau, G, Faas, R. W., Grant, J., et al. 1993. An ecological cascade effect: migratory birds affect stability of intertidal sediments. Limnology and Oceanography, 38: 225231.Google Scholar
Da Cunha Lana, P., and Guiss, C. 1992. Macrofauna-plant-biomass interactions in a euhaline salt marsh in Paranagua Bay (SE Brazil). Marine Ecology Progress Series, 80: 5764.Google Scholar
Darby, F. A., and Turner, R. E. 2008a. Effects of eutrophication on salt marsh root and rhizome biomass accumulation. Marine Ecology Progress Series, 363: 6370.Google Scholar
Darby, F. A., and Turner, R. E. 2008b. Below- and aboveground biomass of Spartina alterniflora: response to nutrient addition in a Louisiana Salt Marsh. Estuaries and Coasts, 31: 326334.Google Scholar
Davenport, T. M., Seitz, R. D., Knick, K. E., and Jackson, N. 2018. Living shorelines support nearshore benthic communities in Upper and Lower Chesapeake Bay. Estuaries and Coasts, 41: 197206.Google Scholar
Davey, E., Wigand, C., Johnson, R., Sundberg, K., Morris, J., Roman, C. T., Davey, E. et al. 2011. Use of computed tomography imaging for quantifying coarse roots, rhizomes, peat, and particle densities in marsh soils. Ecological Applications, 21: 21562171.Google Scholar
Day, J. W. Jr, Kemp, G. P., Reed, D. J., Cahoon, D. R., Boumans, R. M., Suhayda, J. M., and Gambrell, R.. 2011. Vegetation death and rapid loss of surface elevation in two contrasting Mississippi delta salt marshes: the role of sedimentation, autocompaction and sea level rise. Ecological Engineering, 37: 228240.Google Scholar
Deegan, L. A., Johnson, D. S., Warren, R. S., Peterson, B. J., Fleeger, J. W., Fagherazzi, S., and Wollheim, W. 2012. Coastal eutrophication as a driver of salt marsh loss. Nature, 490: 388394.Google Scholar
Delaune, R. D., Baumann, R. H., and Gosselink, J. G. 1983. Relationships among vertical accretion, coastal submergence, and erosion in a Louisiana Gulf Coast marsh. Journal of Sedimentary Research, 53: 147157.Google Scholar
Delaune, R. D., Nyman, J. A., and Patrick, W. H. 1994. Peat collapse, ponding and wetland loss in a rapidly submerging coastal marsh. Journal of Coastal Research, 10: 10211030.Google Scholar
Delaune, R. D., Patrick, W. H., and Buresh, R. J. 1978. Sedimentation rates determined by 137Cs dating in a rapidly accreting salt marsh. Nature, 275: 532533.Google Scholar
Dionne, J.-C. 1985. Tidal marsh erosion by geese, St. Lawrence Estuary, Québec. Géographie Physique et Quaternaire, 39: 99105.Google Scholar
Dupraz, C. and Visscher, P. T. 2005. Microbial lithification in marine stromatolites and hypersaline mats. Trends in Microbiology, 13: 429438.Google Scholar
Dyer, K. R. 1988. Fine sediment particle transport in estuaries. In: Dronkers, J., and van Leussen, W., eds, Physical Processes in Estuaries. Springer, Berlin, Heidelberg, pp. 295310.Google Scholar
Ellison, A. M., Bertness, M. D., and Miller, T. 1986. Seasonal patterns in the belowground biomass of Spartina alterniflora (Gramineae) across a tidal gradient. American Journal of Botany, 73: 15481554.Google Scholar
Escapa, M., Minkoff, D. R., Perillo, G. M. E., and Iribarne, O. 2007. Direct and indirect effects of burrowing crab Chasmagnathus granulatus activities on erosion of Southwest Atlantic Sarcocornia-dominated marshes. Limnology and Oceanography, 52: 23402349.Google Scholar
Escapa, M., Perillo, G. M. E., and Iribarne, O. 2008. Sediment dynamics modulated by burrowing crab activities in contrasting SW Atlantic intertidal habitats. Estuarine, Coastal and Shelf Science, 80: 365373.Google Scholar
Escapa, C. M., Perillo, G. M. E., Iribarne, O. 2015. Biogeomorphically driven salt pan formation in Sarcocornia-dominated salt-marshes. Geomorphology, 228: 147157.Google Scholar
Fagherazzi, S., FitzGerald, D. M., Fulweiler, R. W., Hughes, Z., Wiberg, P. L., McGlathery, K. J., Morris, J. T., Tolhurst, T. J., Deegan, L. A. and Johnson, D. S. 2013a. Ecogeomorphology of salt marshes. Treatise on Geomorphology, 12: 182212.Google Scholar
Fagherazzi, S., Wiberg, P. L., Temmerman, S., Struyf, E., Zhao, Y., and Raymond, P. A. 2013b. Fluxes of water, sediments, and biogeochemical compounds in salt marshes. Ecological Processes, 2: 116.Google Scholar
Fanjul, E., Grela, M. A., and Iribarne, O. 2007. Effects of the dominant SW Atlantic intertidal burrowing crab Chasmagnathus granulatus on sediment chemistry and nutrient distribution. Marine Ecology Progress Series, 341: 177190.Google Scholar
Farron, S. J. 2018. Morphodynamic responses of salt marshes to sea-level rise: upland expansion, drainage evolution, and biological feedbacks. PhD Dissertation, Boston University, 159 p.Google Scholar
FitzGerald, Duncan M., Fenster, M.S., Argow, B.A., and Buynevich, I.V. 2008. Coastal impacts due to sea-level rise. Annual Review of Earth and Planetary Sciences, 36: 601647.Google Scholar
Ford, Mark A., and Grace, James B.. 1998. Effects of vertebrate herbivores on soil processes, plant biomass, litter accumulation and soil elevation changes in a coastal marsh. Journal of Ecology, 86: 974982.Google Scholar
French, J. R., and Spencer, T. 1993. Dynamics of sedimentation in a tide-dominated backbarrier salt marsh, Norfolk, UK. Marine Geology, 110: 315331.Google Scholar
French, J. R., and Stoddart, D. R. 1992. Hydrodynamics of Salt Marsh Creek Systems: Implications for Marsh Morphological Development and Material Exchange. Earth Surface Processes and Landforms 17: 235252.Google Scholar
Friedrichs, C. T., and Perry, J. E. 2001. Tidal salt marsh morphodynamics: A synthesis. Journal of Coastal Research, Special issue, no. 27: 7–37.Google Scholar
Gauthier, G., Giroux, J. F., Reed, , Béchet, A., and Bélanger, L. 2005. Interactions between land use, habitat use, and population increase in Greater Snow Geese: What are the consequences for natural wetlands? Global Change Biology, 11: 856868.Google Scholar
Gedan, K. B., Kirwan, M. L., Wolanski, E., Barbier, E. B., and Silliman, B. R. 2011. The present and future role of coastal wetland vegetation in protecting shorelines: answering recent challenges to the paradigm. Climatic Change, 106: 729.Google Scholar
Genoni, G. P. 1991. Increased burrowing by fiddler crabs Uca rapax (Smith) (Decapoda : Ocypodidae) in response to low food supply. Journal of Experimental Marine Biology and Ecology, 147: 267285.Google Scholar
Giosan, L., Syvitski, J., Constantinescu, S., and Day, J. 2014. Climate change: protect the world’s deltas. Nature, 516: 3133.Google Scholar
Gleason, M. L., Elmer, D. A., Pien, N. C., Fisher, J. S. 1979. Effects of stem density upon sediment retention by salt marsh cord grass, Spartina alterniflora Loisel. Estuaries, 2: 271273.Google Scholar
Grant, J., Bathmann, U. V., and Mills, E. L. 1986. The interaction between benthic diatom films and sediment transport. Estuarine Coastal Shelf Science, 23: 225238.Google Scholar
Gribsholt, B., Kostka, J. E., and Kristensen, E. 2003. Impact of fiddler crabs and plant roots on sediment biogeochemistry in a Georgia saltmarsh. Marine Ecology Progress Series, 259: 237251.Google Scholar
Gross, M. F., Hardisky, M. A., Wolf, P. L., and Klemas, V. 1991. Relationship between aboveground and belowground biomass of Spartina alterniflora (Smooth Cordgrass). Estuaries, 14: 180191.Google Scholar
Gutiérrez, J. L., and Iribarne, O. 1999. Role of Holocene beds of the stout razor clam Tagelus plebeius in structuring present benthic communities. Marine Ecology Progress Series, 185: 213228.Google Scholar
Gutiérrez, J. L., Jones, C. G., Groffman, P. M., Findlay, S. E. G, Iribarne, O., Ribeiro, P. D., and Bruschetti, C. M. 2006. The contribution of crab burrow excavation to carbon availability in surficial salt-marsh sediments. Ecosystems, 9: 647658.Google Scholar
Gutiérrez, J. L., Jones, C. G., Strayer, D. L., and Iribarne, O. 2003. Mollusks as ecosystem engineers : The role of shell production in aquatic habitats. Oikos, 101: 7990.Google Scholar
Hannaford, J.Pinn, E. H., and Diaz, A. 2006. The impact of sika deer grazing on the vegetation and infauna of Arne saltmarsh. Marine Pollution Bulletin, 53: 5662.Google Scholar
Hatton, R. S., DeLaune, R. D., and Patrick, W. H. Jr. 1983. Sedimentation, accretion, and subsidence in marshes of Barataria Basin, Louisiana. Limnology and Oceanography, 28: 494502.Google Scholar
Hazelden, J., and Boorman, L. A. 2001. Soils and “managed retreat” in South East England. Soil Use and Management, 17: 150154.Google Scholar
Holdredge, C., Bertness, M. D., and Altieri, A. H. 2008. Role of crab herbivory in die-off of new england salt marshes. Conservation Biology, 23: 672679.Google Scholar
Holm, G. O. 2006. Nutrient constraints on plant community production and organic matter accumulation of subtropical floating marshes. PhD dissertation, Louisiana State University, Baton Rouge, Louisiana.Google Scholar
Hopkinson, C. S., Gosselink, J. G., and Parrondo, R. T. 1980. Production of coastal louisiana marsh plants calculated from phenometric techniques Ecology, 61: 10911098.Google Scholar
Howe, A. J., Rodriguez, J. F., and Saco, P. M. 2009. Surface evolution and carbon sequestration in disturbed and undisturbed wetland soils of the Hunter estuary, southeast Australia. Estuarine Coastal Shelf Science, 84: 7583.Google Scholar
Howes, B. L., Goehringer, D. D., and Macey, J. W. H. 1986 Factors controlling the growth form of Spartina alterniflora: feedbacks between above-ground production, sediment oxidation, nitrogen and salinity. Journal of Ecology, 74: 881898.Google Scholar
Howes, B. L., Howarth, R. W., Teal, J. M., and Valiela, I. 1981. Oxidation-reduction potentials in a saltmarsh: spatial patterns and interactions with primary production. Limnology and Oceanography, 26: 350360.Google Scholar
Howes, N. C., FitzGerald, D. M., Hughes, Z. J., Georgiou, I. Y., Kulp, M. A., Miner, M. D., Smith, J. M., and Barras, J. A. 2010. Hurricane-induced failure of low salinity wetlands. Proceedings of the National Academy of Sciences of the USA, 107: 1401414019.Google Scholar
Hu, K., Chen, Q., and Wang, H. 2015. A numerical study of vegetation impact on reducing storm surge by wetlands in a semi-enclosed estuary. Coastal Engineering, 95: 6676.Google Scholar
Hughes, Z. J., FitzGerald, D. M., Wilson, C. A., Pennings, S. C. Wiçski, K., and Mahadevan, A. 2009. Rapid headward erosion of marsh creeks in response to relative sea level rise. Geophysical Research Letters, 36: 15.Google Scholar
Hyun, J., Smith, A. C., and Kostka, J. E. 2007. Relative contributions of sulfate- and iron(III) reduction to organic matter mineralization and process controls in contrasting habitats of the Georgia saltmarsh. Applied Geochemistry, 22: 26372651.Google Scholar
Iribarne, O., Bortolus, A., Botto, F. 1997. Between-habitat differences in burrow characteristics and trophic modes in the south western Atlantic burrowing crab Chasmagnathus granulata. Marine Ecology Progress Series, 155: 137145.Google Scholar
Jaramillo, E., and Lunecke, K. 1988. The role of sediments in the distribution of Uca pugilator (Bosc) and Uca pugnax (Smith) (Crustacea, Brachyura) in a salt marsh at Cape Cod. Meeresforschung, 32: 4652.Google Scholar
Jones, C. G., Lawton, J. H., and Shachak, M. 1994. Organisms as ecosystem engineers. Oikos 69: 373386.Google Scholar
Jordan, T., and Valiela, I. 1982. A nitrogen budget of the ribbed mussel, Geukensia demissa, and its significance in nitrogen flow in a New England salt marsh. Limnology and Oceanography, 27: 7590.Google Scholar
Julien, A. 2018. Quantifying the demographics, habitat characteristics, and foundation species role of the ribbed mussel (Geukensia demissa) in South Carolina salt marshes. Masters thesis, Dept of Biology, College of Charleston, South Carolina.Google Scholar
Katrak, G., Dittmann, S., and Seurant, L. 2008. Spatial variation in burrow morphology of the mud shore crab Helograpsus haswellianus (Brachyura, Grapsidae) in South Australian saltmarshes. Marine and Freshwater Research, 59, 902911.Google Scholar
Katz, L. C. 1980. Effects of burrowing by the fiddler crab, Uca pugnax (Smith). Estuarine Coastal and Marine Science, 11: 233237.Google Scholar
Kaye, C. A., and Barghoorn, E. S. 1964. Late Quaternary sea-level change and crustal rise at Boston Massachusetts, with notes on autocompaction of peat. GSA Bulletin, 75: 6380.Google Scholar
Kennish, M. J. 2001. Coastal salt marsh systems in the US: a review of anthropogenic impacts. Journal of Coastal Research, 17: 731748.Google Scholar
Kesel, R. H., Yodis, E. G., and McCraw, D. J. 1992. An approximation of the sediment budget of the Lower Mississippi River prior to major human modification. Earth Surface Processes and Landforms, 17: 711722.Google Scholar
Keusenkothen, M. A. 2002. The effects of deer trampling in a salt marsh. MS thesis, East Carolina University.Google Scholar
Keusenkothen, M. A., and Christian, R. R. 2004. Responses of salt marshes to disturbance in an ecogeomorphological context, with a case study of trampling by deer. In: Fagherazzi, S., Marani, M., and Blum, L., eds., The Ecogeomorphology of Tidal Marshes, Volume 59. John Wiley, Hoboken, NJ, pp. 203230.Google Scholar
King, G. M., Klug, M. J., Wiegert, R. G., and Chalmers, A. G. 1982. Relation of soil water movement and sulfide concentration of Spartina alterniflora production in a Georgia Salt Marsh. Science, 218: 6163.Google Scholar
Kirchner, J. W., Dietrich, W. E., Iseya, F., and Ikeda, H. 1990. The variability of critical shear stress, friction angle, and grain protrusion in water-worked sediments. Sedimentology, 37: 647672.Google Scholar
Kirwan, M. L., and Guntenspergen, G. R. 2010. The influence of tidal range on the stability of coastal marshland. Journal of Geophysical Research, 115: F02009, doi:10.1029/2009JF001400.Google Scholar
Kirwan, M. L., Guntenspergen, G. R., D’Alpaos, A., Morris, J., Mudd, S. M., and Temmerman, S. 2010. Limits on the adaptability of coastal marshes to rising sea level. Geophysical Research Letters, 37(23). https://doi.org/10.1029/2010GL045489Google Scholar
Kirwan, M. L., and Megonigal, J. P. 2013. Tidal wetland stability in the face of human impacts and sea-level rise. Nature, 504: 5360.Google Scholar
Kirwan, M. L., and Temmerman, S. 2009. Coastal marsh response to historical and future sea-level acceleration. Quaternary Science Reviews, 28: 18011808.Google Scholar
Kirwan, M. L., Temmerman, S., Skeehan, E. E., Guntenspergen, G. R., and Fagherazzi, S. 2016. Overestimation of marsh vulnerability to sea level rise. Nature Climate Change, 6: 253260.Google Scholar
Knutson, P. L., Brochu, R. A., and See, W. N. 1982. Wave damping in Spartina alterniflora marshes. Wetlands, 2: 87104.Google Scholar
Kobayashi, N., Raichle, A., and Asano, T. 1993. Wave attenuation by vegetation. Journal of Waterway, Port, Coastal, and Ocean Engineering Technical Report, 119(1). https://doi.org/10.1061/(ASCE)0733-950X(1993)119:1(30)Google Scholar
Koch, E. W., Barbier, E. D., Silliman, B. R., Reed, D. J., Perillo, G. M. E., Hacker, S. D., Granek, E. F., et al. 2009. Non-linearity in ecosystem services: temporal and spatial variability in coastal protection. Frontiers in Ecology and the Environment, 7: 2937.Google Scholar
Kokot, R. R. 2004. Erosión en la costa patagónica por cambio climático. Revista de la Asociación Geológica Argentina, 59: 715726.Google Scholar
Koo, B. J., Kwon, K. K., and Hyun, J. H. 2005. The sediment-water interface increment due to the complex burrows of macrofauna in a tidal flat. Ocean Science Journal, 40: 221227.Google Scholar
Kostka, J. E., Gribsholt, B., Petrie, E., Dalton, D., Skelton, H., and Kristensen, E. 2002. The rates and pathways of carbon oxidation in bioturbated saltmarsh sediments. Limnology and Oceanography, 47: 230240.Google Scholar
Kraeuter, J. N. 1976. Biodeposition by salt-marsh invertebrates. Marine Biology, 35: 215223.Google Scholar
Laegdsgaard, P. 2006. Ecology, disturbance and restoration of coastal saltmarsh in Australia: A Review. Wetlands Ecology and Management, 14: 379399.Google Scholar
Langley, J. A., McKee, K. L., Cahoon, D. R., Cherry, J. A. and Megonigal, P. 2009. Elevated CO2 stimulates marsh elevation gain, counterbalancing sea-level rise. Proceedings of the National Academy of Sciences of the USA, 106: 61826186.Google Scholar
Lanuru, M. 2008. Measuring critical erosion shear stress of intertidal sediments with EROMES erosion device. Torani Journal of Marine Science and Fisheries, 18: 390397.Google Scholar
Leeder, M. R. 1998. Lyell’s Principles of Geology: Foundations of sedimentology. Geological Society of London, Special Publication, 143: 95110.Google Scholar
Leonard, L. A., and Luther, M. E. 1995. Flow hydrodynamics in tidal marsh cano-pies. Limnology and Oceanography 18: 14741484.Google Scholar
Letzsch, W. S., and Frey, R. W. 1980. Deposition and erosion in a Holocene salt marsh, Sapelo Island, Georgia. Journal of Sedimentary Research, 50: 529542.Google Scholar
Li, H., and Yang, S. L. 2009. Trapping effect of tidal marsh vegetation on suspended sediment, Yangtze Delta. Journal of Coastal Research, 25: 915924.Google Scholar
Lightbody, A. F., and Nepf, H. M. 2006. Prediction of velocity profiles and longitudinal dispersion in salt marsh vegetation. Limnology and Oceanography, 51: 218228.Google Scholar
Lynch, J. J., O’Neil, E., and Lay, D. W. 1947. Management significance of damage by geese and muskrats to Gulf Coast marshes. Journal of Wildlife Management, 11: 5076.Google Scholar
Madsen, K. N., Nilsson, P., and Sunback, K. 1993. The influence of benthic microalgae on the stability of a subtidal sediment. Journal of Experimental Marine Biology and Ecology, 170: 159177.Google Scholar
Madsen, J. D., Chambers, P. A., James, W. F., Koch, E. W., and Westlake, D. F. 2001. The interaction between water movement, sediment dynamics and submersed macrophytes. Hydrobiologia, 444: 7184.Google Scholar
Marani, M., D’Alpaos, A., Lanzoni, S., Carniello, L., and Rinaldo, A. 2007. Biologically-controlled multiple equilibria of tidal landforms and the fate of the Venice lagoon. Geophysical Research Letters, 34: 15.Google Scholar
Mariotti, G., and Carr, J. 2014. Dual role of salt marsh retreat: Long‐term loss and short‐term resilience. Water Resources Research, 50: 29632974.Google Scholar
Mariotti, G., and Fagherazzi, S. 2010. A numerical model for the coupled long‐term evolution of salt marshes and tidal flats. Journal of Geophysical Research Earth Surface, 115 (F1). https://doi.org/10.1029/2009JF00132.Google Scholar
Mariotti, G., and Fagherazzi, S. 2013. A two-point dynamic model for the coupled evolution of channels and tidal flats. Journal of Geophysical Research, 118: 13871399.Google Scholar
Mariotti, G., Kearney, W., and Fagherazzi, S. 2016. Soil creep in salt marshes. Geology, 44: 459462.Google Scholar
McCraith, B. J., Gardner, L. R., Wethey, D. S., and Moore, W. S. 2003. The effect of fiddler crab burrowing on sediment mixing and radionuclide profiles along a topographic gradient in a southeastern salt marsh. Journal of Marine Research, 61: 359390.Google Scholar
Mehta, A. J. 1996. Interaction between fluid mud and water waves. In: Singh, V. P., and Hager, W. H., Environmental Hydraulics. Kluwer, Dordretcht, pp. 153187.Google Scholar
Melo, W. D., Perillo, G. M. E., Perillo, M. M., Schilizzi, R., and Piccolo, M. C. 2013. Late Pleistocene-Holocene deltas in the southern Buenos Aires Province, Argentina. In: Young, G., and Perillo, G. M. E., eds., Deltas: Landforms, Ecosystems and Human Activities. IAHS Press, Wallingford, UK. 358: 187195.Google Scholar
Melo, W. D., Schillizzi, R., Perillo, G. M. E. y Piccolo, M. C. 2003. Influencia del área continental pampeana sobre el origen y la morfología del estuario de Bahía Blanca. Revista de la Asociación Argentina de Sedimentología, 10: 6572.Google Scholar
Mendelssohn, I. A., McKee, K. L., and Patrick, W. H. Jr 1981. Oxygen deficiency in Spartina alterniflora roots: metabolic adaptation to anoxia. Science, 214: 439441.Google Scholar
Mendelssohn, I. A., and Seneca, E. D. 1980. The influence of soil drainage on the growth of salt marsh cordgrass Spartina alterniflora in North Carolina. Estuarine and Coastal Marine Science, 11: 2740.Google Scholar
Meyer, D. L., Townsend, E. C., and Thayer, G. W. 1997. Stabilization and erosion control value of oyster cultch for intertidal marsh. Restoration Ecology, 5: 9399.Google Scholar
Millette, T., Argow, B., Marcano, E., Hayward, C., Hopkinson, C., and Valentine, V. 2010. Salt marsh geomorphological analyses via integration of multitemporal multispectralremote sensing with LIDAR and GIS. Journal of Coastal Research, 26: 809816.Google Scholar
Minkoff, D. R., Escapa, M., Ferramola, F. E., Maraschín, S. D., Pierini, J. O., Perillo, G. M. E., and Delrieux, C. 2006. Effects of crab–halophytic plant interactions on creek growth in a S. W. Atlantic salt marsh: a cellular automata model. Estuarine, Coastal and Shelf Science, 69: 403413.Google Scholar
Minkoff, D. R., Escapa, C. M., Ferramola, F. E., and Perillo, G. M. E. 2005. Erosive processes due to physical – biological interactions based in a cellular automata model. Latin American Journal of Sedimentology and Basin Analysis, 12: 2534.Google Scholar
Mitsch, W. J., and Gosselink, J. G. 2000. Wetlands. 3rd edition. John Wiley, New York.Google Scholar
Molina, L. M., Valiñas, M. S., Pratolongo, P., Elias, R., and Perillo, G. M. E. 2017. Effect of Micropogonias furnieri on the stability of the sediment of salt marshes – an issue to be resolved. Estuaries and Coasts, 40: 17951807.Google Scholar
Moller, I., Kudella, M., Rupprecht, F., Spencer, T., Paul, M., van Wesenbeeck, B. K., Wolters, G., et al. 2014. Wave attenuation over coastal salt marshes under storm surge conditions. Nature Geoscience, 7: 727731.Google Scholar
Moller, I., and Spencer, T. 2002. Wave dissipation over macro-tidal saltmarshes: Effects of marsh edge typology and vegetation change. Journal of Coastal Research: Special Issue 36 – International Coastal Symposium (ICS 2002): 506521.Google Scholar
Montague, C. L. 1980. A natural history of temperate Western Atlantic fiddler crabs (Genus Uca) with reference to their impact on the salt marsh. Contributions in Marine Science, 23: 2555.Google Scholar
Montague, C. L. 1982. The influence of fiddler crab burrowing on metabolic processes in saltmarsh sediments. In: Kennedy, V. S., ed., Estuarine Comparisons. Academic Press, San Francisco, pp. 283301.Google Scholar
Montgomery, D. R., and Dietrich, W. E. 1988. Where do channels begin? Nature, 336: 232234.Google Scholar
Morris, J. M., Sundareshwar, P. V., Nietch, C. T., Kjerfve, B., and Cahoon, D. R. 2002. Responses of coastal wetlands to rising sea level. Ecology, 83: 28692877.Google Scholar
Mouton, E C., and Felder, D. L. 1996. Burrow distributions and population estimates for the fiddler crabs Uca spinicarpa and Uca longisignalis in a Gulf of Mexico salt marsh. Estuaries and Coasts, 19: 5161.Google Scholar
Murray, J. M. H., Meadows, A., and Meadows, P. S. 2002. Biogeomorphological implications of microscale interactions between sediment geotechnics and marine benthos: A review. Geomorphology, 47: 1530.Google Scholar
Nelson, J., Wilson, R., Coleman, F., Koenig, C., DeVries, D., Gardner, C., and Chanton, J. 2012. Flux by fin: fish-mediated carbon and nutrient flux in the northeastern Gulf of Mexico. Marine Biology, 159: 365372.Google Scholar
Nepf, H. M. 1999. Drag, turbulence, and diffusion in flow through emergent vegetation. Water Resources Research, 35: 479489.Google Scholar
Neumeier, U., and Amos, C. L. 2006. The influence of vegetation on turbulence and flow velocities in European salt‐marshes. Sedimentology, 53: 259277.Google Scholar
Neumeier, U., and Ciavola, P. 2004. Flow resistance and associated sedimentary processes in a Spartina maritima salt-marsh. Journal of Coastal Research, 20: 435447.Google Scholar
Niering, W., and Warren, R. S. 1980. Vegetation patterns and processes in New England Salt Marshes. BioScience, 30: 301307.Google Scholar
NOAA. 2018. Green Infrastructure Effectiveness Database: https://coast.noaa.gov/digitalcoast/training/gi-database.htmlGoogle Scholar
Nyman, J. A., Crozier, C., and DeLaune, R. D. 1995a. Roles and patterns of hurricane sedimentation in an estuarine marsh landscape. Estuaries, Coastal and Shelf Science, 40: 665679.Google Scholar
Nyman, J. A., Delaune, R. D., Patrick, W. H. Jr. 1990. Wetland soil formation in the rapidly subsiding Mississippi River Deltaic Plain: mineral and organic matter relationships. Estuarine, Coastal and Shelf Science, 31: 5769.Google Scholar
Nyman, J. A., DeLaune, R. D., Pezeshki, S. R., and Patrick, W. H. Jr. 1995b. Organic matter cycling and marsh stability in a rapidly submerging estuarine marsh. Estuaries, 18: 207218.Google Scholar
Nyman, J. A., Walters, R., Delaune, R. D., Patrick, W. H. Jr. 2006. Marsh vertical accretion via vegetative growth. Estuaries Coastal and Shelf Science, 69: 370380.Google Scholar
O’Donnell, J. E. D. 2017. Living shorelines: A review of literature relevant to New England coasts. Journal of Coastal Research, 332: 435451.Google Scholar
Odum, W. 1988. Comparative ecology of tidal freshwater and salt marshes. Annual Review of Ecology and Systematics, 19: 147176.Google Scholar
Onorevole, K. M., Thompson, S. P., and Piehler, M. F. 2018. Living shorelines enhance nitrogen removal capacity over time. Ecological Engineering, 120: 238248.Google Scholar
Pan, J., Bournod, C. N., Pizani, N. V., Cuadrado, D. G., and Carmona, N. B. 2013. Characterization of microbial mats from a siliciclastic tidal flat (Bahía Blanca Estuary, Argentina). Geomicrobiology Journal, 30: 665674.Google Scholar
Pennings, S. C., and Callaway, R. 1992. Salt marsh plant zonation: the relative importance of competition and physical factors. Ecology, 73: 681690.Google Scholar
Pennings, S. C., Carefoot, T. H., Siska, E. L., Chase, M.G., and Page, T. A. 1998. Feeding preferences of a generalist salt-marsh crab: relative importance of multiple plant traits. Ecology, 79: 19681979.Google Scholar
Pennings, S. C., and Silliman, B. R. 2005. Linking biogeography and community ecology : Latitudinal variation in plant-herbivore interaction strength. Ecology, 86: 23102319.Google Scholar
Perillo, G. M. E. 2019. Geomorphology of tidal courses and depressions. In: Perillo, G. M. E., Wolanski, E., Cahoon, D. R., and Hopkinson, C., eds., Coastal Wetlands: An Integrated Ecosystem Approach. Elsevier, Amsterdam, pp. 185210.Google Scholar
Perillo, G. M. E., Drapeau, G., Piccolo, M. C., and Chaouq, N. 1993. Tidal circulation pattern on a tidal flat, Minas Basin, Canada. Marine Geology, 112: 219236Google Scholar
Perillo, G. M. E., and Iribarne, O. 2003a. New mechanisms studied for creek formation in tidal flats: from crabs to tidal channels. EOS American Geophysical Union Transactions, 84: 15.Google Scholar
Perillo, G. M. E., and Iribarne, O. 2003b. Processes of tidal channels develop in salt and freshwater marshes. Earth Surface Processes and Landforms, 28: 14731482.Google Scholar
Perillo, G. M. E., Minkoff, D. R., and Piccolo, M. C. 2005. Novel mechanism of stream formation in coastal wetlands by crab–fish–groundwater interaction. Geo-Marine Letters, 25: 214220.Google Scholar
Pestrong, R. 1965. The development of drainage patterns on tidal marshes. Stanford University Publications. Earth Science, 10(2): 1–87.Google Scholar
Pestrong, R. 1972. Tidal-flat sedimentation at cooley landing, Southwest San Francisco bay. Sedimentary Geology, 8: 251288.Google Scholar
Peterson, G. W., and Turner, R. E. 1994. The value of salt marsh edge vs interior as a habitat for fish and decapod crustaceans in a Louisiana tidal marsh. Estuaries, 17: 235262.Google Scholar
Pethick, J. S. 1980. Velocity surges and asymmetry in tidal channels. Estuarine Coastal and Marine Science, 11: 331345.Google Scholar
Pethick, J. S. 1992. Saltmarsh geomorphology. In: Allen, J. R. L., and Pye, K., eds, Saltmarshes: Morphodynamics, Conservation and Engineering Significance, Cambridge University Press, Cambridge, UK. pp. 4162.Google Scholar
Piazza, B. P., Banks, P. D., and La Peyre, M. K. 2005. The potential for created oyster shell reefs as a sustainable shoreline protection strategy in Louisiana. Restoration Ecology, 13: 499506.Google Scholar
Postma, H. 1961. Transport and accumulation of suspended matter in the Dutch Wadden Sea. Netherlands Journal of Sea Research, 1: 148180.Google Scholar
Pratolongo, P. D., Mazzon, C., Zapperi, G., Piovan, M. J., and Brinson, M. M. 2013. Land cover changes in tidal salt marshes of the Bahía Blanca estuary (Argentina) during the past 40 years. Estuarine, Coastal and Shelf Science, 133: 2331.Google Scholar
Pratolongo, P. D., Perillo, G. M. E., and Piccolo, M. C. 2010. Combined effects of waves and marsh plants on mud deposition events at a mudflat-saltmarsh edge. Estuarine, Coastal and Shelf Sciences, 87: 207212.Google Scholar
Priestas, A. M., and Fagherazzi, S. 2011. Morphology and hydrodynamics of wave-cut gullies. Geomorphology, 131: 113.Google Scholar
Pye, K., and French, P. W. 1993. Erosion and Accretion Processes on British Saltmarshes. Volume One. Introduction: Saltmarsh Processes and Morphology. Report No. ES19. Ministry of Agriculture, Fisheries and Food. Cambridge Environmental Research Consultants, Cambridge.Google Scholar
Reddy, K. R., and Delaune, R. D. 2008. Biogeochemistry of Wetlands: Science and Applications. CRC Press, Boca Raton, FL.Google Scholar
Redfield, A. C. 1965. Ontogeny of a saltmarsh estuary. Science, 147: 5055.Google Scholar
Redfield, A. C. 1972. Development of a New England Salt Marsh. Ecological Monographs, 42: 201237.Google Scholar
Redfield, A. C., and Rubin, M. 1962. The age of salt marsh peat and its relation to recent changes in sea level at Barnstable, Massachusetts. Proceeding of the National Academy of Science of the United States of America, 48: 17281735.Google Scholar
Reed, D. J. 1988. Sediment dynamics and deposition in a retreating coastal salt marsh. Estuarine, Coastal and Shelf Science, 26: 6769.Google Scholar
Reed, D. J., Spencer, T., Murray, A., French, J. R., and Leonard, L. 1999. Marsh surface sediment deposition and the role of tidal creeks: implications for created and managed coastal marshes. Journal of Coastal Conservation, 5: 8190.Google Scholar
Rice, D. L. 1986. Early diagenesis in bioadvective sediments: Relationships between the diagenesis of beryllium-7, sediment reworking rates, and the abundance of conveyor-belt deposit-feeders. Journal of Marine Research, 44: 149184.Google Scholar
Ringold, P. 1979. Burrowing, root mat density, and the distribution of fiddler crabs in the Eastern United States. Journal of Experimental Marine Biology and Ecology, 36: 1121.Google Scholar
Robertson, T. L., and Weis, J. S. 2005. A comparison of epifaunal communities associated with the stems of salt marsh grasses Phragmites australis and Spartina alterniflora. Wetlands, 25: 17.Google Scholar
Roman, C. T., Peck, J. A., Allen, J. R., King, J. W., and Appleby, P. G. 1997. Accretion of a New England (U.S.A.) salt marsh in response to inlet migration, storms, and sea-level rise. Estuarine Coastal and Shelf Science, 46: 717727.Google Scholar
Schwimmer, R. 2001, Rates and processes of marsh shoreline erosion in Rehoboth Bay, Delaware, U.S.A. Journal of Coastal Research, 17: 672683.Google Scholar
Schwimmer, R., and Pizzuto, J. 2000. A model for the evolution of marsh shorelines. Journal of Sedimentary Research, 70: 10261035.Google Scholar
Scordo, F., Bohn, V., Piccolo, M. C., and Perillo, G. M. 2018. Mapping and monitoring lakes intra-annual variability in semi-arid regions: a case study in Patagonian Plains (Argentina). Water, 10: 889.Google Scholar
Scyphers, S. B., Powers, S. P., Heck, K. L., and Byron, D. 2011. Oyster reefs as natural breakwaters mitigate shoreline loss and facilitate fisheries. PLOS ONE. 6(8). doi:10.1371/journal.pone.0022396.Google Scholar
Shaffer, G. P., Sasser, C. E., Gosselink, J. G., and Rejmanek, M. 1992. Vegetation dynamics in the emerging Atchafalaya Delta, Louisiana, USA. Journal of Ecology, 80: 677687.Google Scholar
Sharma, P., Gardner, L. R., Moore, W. S., and Bollinger, M. S. 1987. Sedimentation and bioturbation in a salt marsh as revealed by 210Pb, 137Cs, and 7Be studies. Limnology and Oceanography, 32: 313326.Google Scholar
Shepard, C. C., Crain, C. M., and Beck, M. W. 2011. The protective role of coastal marshes: a systematic review and meta-analysis. PLOS ONE, 6(11): e27374. https://doi.org/10.1371/journal.pone.0027374Google Scholar
Shi, B. W., Yang, S. L., Wang, Y. P., Bouma, T. J., and Zhu, Q. 2012. Relating accretion and erosion at an exposed tidal wetland to the bottom shear stress of combined current-wave action. Geomorphology, 138: 380389.Google Scholar
Silliman, B. R., Van De Koppel, J., Bertness, M. D., Stanton, L. E., and Mendelssohn, I. A. 2005. Drought, snails, and large-scale die-off of southern U.S. salt marshes. Science, 310: 18031807.Google Scholar
Silliman, B. R., and Zieman, J. 2001. Top-down control of Spartina alterniflora production by periwinkle grazing in a Virginia salt marsh. Ecology, 82: 28302845.Google Scholar
Slatyer, R. A., Fok, E. S. Y., Hocking, R., and Backwell, P. R.Y. 2008. Why do fiddler crabs build chimneys? Biology Letters of the Royal Society, 4: 616618.Google Scholar
Smith, S. M. 2009. Multi-decadal changes in salt marshes of Cape Cod, Massachusetts: a photographic analysis of vegetation loss, species shifts, and geomorphic change. Northeastern Naturalist, 16: 183208.Google Scholar
Smith, J. E., Bentley, S. J., Snedden, G. A., and White, C. 2015. What role do hurricanes play in sediment delivery to subsiding river deltas? Scientific Reports, 5, Article number: 17582.Google Scholar
Smith, J. M., and Frey, R. W. 1985. Biodeposition by the ribbed mussel Geukensia demissa in a salt marsh, Sapelo Island, Georgia. Journal of Sedimentary Research, 55: 817828.Google Scholar
Soudry, D. 2000. Microbial phosphate sediment. In: Riding, R. E., and Awramik, S. M., eds., Microbial Sediments. Springer-Verlag, Berlin, pp. 127136.Google Scholar
Spalding, M. D., Ruffo, S., Lacambra, C., Meliane, I., Zeitlin Hale, L., Shepard, C. C., and Beck, M. W. 2014. The role of ecosystems in coastal protection: Adapting to climate change and coastal hazards. Ocean and Coastal Management, 90: 5057.Google Scholar
Steel, T. J., and Pye, K. 1997. The development of saltmarsh tidal creek networks: Evidence from the U.K. Proceedings of the Canadian Coastal Conference, pp. 267–280.Google Scholar
Stevenson, J. C., Ward, L. G., and Kearney, M. S. 1986. Vertical accretion in marshes with varying rates of sea-level rise. In: Wolfe, D. A., ed.,  Estuarine Variability, Academic Press, New York, pp. 241259.Google Scholar
Stolz, J. F. 2000. Structure of microbial mats and biofilms. In: Riding, R. E., and Awramik, S. M., eds., Microbial Sediments. Springer-Verlag, Berlin, pp. 18.Google Scholar
Stumpf, R. P. 1983. The process of sedimentation on the surface of a salt marsh. Estuarine, Coastal and Shelf Science, 17: 495508.Google Scholar
Syvitsky, J., Kettner, A. J., Overeem, I., Hutton, E. W., Hannon, M. T., Brakenridge, G. R., Day, J., et al. 2009. Sinking deltas due to human activities. Nature Geoscience, 2: 681686.Google Scholar
Takeda, S., and Kurihara, Y. 1987. The effects of burrowing of Helice tridens (De Haan) on the soil of a salt-marsh habitat. Journal of Experimental Marine Biology and Ecology, 113: 7989.Google Scholar
Taylor, K. L., and Grace, J. B. 1995. The effects of vertebrate herbivory on plant community structure in the coastal marshes of the Pearl River, Louisiana, USA. Wetlands, 15: 6873.CrossRefGoogle Scholar
Teal, J. M. 1958. Distribution of fiddler crabs in Georgia salt marshes. Ecology, 39: 186193.Google Scholar
Temmerman, S., Bouma, T. J., Govers, G., Wang, Z. B., de Vries, M. B., and Herman, P. M. J. 2005. Impact of vegetation on flow routing and sedimentation patterns: Three dimensional modelling for a tidal marsh. Journal of Geophysical Research, 110: F04019, doi: 10.1029/2005JF000301.Google Scholar
Temmerman, S., Bouma, T. J., Van de Koppel, J., Van der Wal, , De Vries, D. M. B., and Herman, P. M. J. 2007. Vegetation causes channel erosion in a tidal landscape. Geology, 35: 631634.Google Scholar
Temmerman, S., Govers, G., Meire, P., and Wartel, S. 2003. Modelling long-term tidal marsh growth under changing tidal conditions and suspended sediment concentrations, Scheldt Estuary, Belgium. Marine Geology, 193: 151169.Google Scholar
Temmerman, S., Govers, G., Wartel, S. and Meire, P. 2004. Modelling estuarine variations in tidal marsh sedimentation: response to changing sea levels and suspended sediment concentrations. Marine Geology, 212: 119.Google Scholar
Thomas, C. R., and Blum, L. K. 2010. Importance of the fiddler crab Uca pugnax to salt marsh soil organic matter accumulation. Marine Ecology Progress Series, 414: 167177.Google Scholar
Tolhurst, T. J., Black, K. S., Shayler, S. A., Mather, S., Black, I., Baker, K., and Paterson, D. M. 1999. Measuring the in Situ erosion shear stress of intertidal sediments with the cohesive strength meter (CSM). Estuarine, Coastal and Shelf Science, 49: 281294.Google Scholar
Turner, R. E. 2010. Doubt and the values of an ignorance-based world view for wetland restoration: Coastal LouisianaEstuaries and Coasts32: 10541068.CrossRefGoogle Scholar
Turner, R. E. 2011. Beneath the salt marsh canopy: loss of soil strength with increasing nutrient loads. Estuaries and Coasts, 34: 10841093.Google Scholar
Turner, R. E., Swenson, E. M., and Milan, C. S. 2002. Organic and inorganic contributions to vertical accretion in salt marsh sediments. In: Weinstein, M. P., and Kreeger, D. A., eds., Concepts and Controversies in Tidal Marsh Ecology. Springer, Dordrecht, pp. 583595.Google Scholar
Underwood, G. J. C, Paterson, D. M., and Parkes, R. J. 1995. The measurement of microbial carbohydrate exopolymers from intertidal sediments. Limnology and Oceanography, 40: 12431453.Google Scholar
Valentine, K., and Mariotti, G. 2019. Wind-driven water level fluctuations drive marsh edge erosion variability in microtidal coastal bays. Continental Shelf Research, 176: 7689.Google Scholar
Valiela, I., Teal, J. M., and Persson, N. Y. 1976. Production and dynamics of experimentally enriched salt marsh vegetation: belowground biomass. Limnology and Oceanography, 21: 245252.Google Scholar
van Asselen, S., Stouthamer, E., and van Asch, Th. W. J. 2009. Effects of peat compaction on delta evolution: a review on processes, responses, measuring and modeling. Earth-Science Reviews, 92: 3551.Google Scholar
van Eerdt, M. 1986. The influence of basic soil and vegetation parameters on salt marsh cliff strength. In: Gardiner, V., ed., International Geomorphology, Part 1, Wiley, Chichester, pp. 10731086.Google Scholar
van Proosdij, D., Davidson-Arnott, R. G. D., and Ollerhead, J. 2006. Controls on spatial patterns of sediment deposition across a macro-tidal salt marsh surface over single tidal cycles. Estuarine, Coastal and Shelf Science, 69: 6486Google Scholar
van Rijn, L. C., van Rossum, H., and Termes, P. 1990. Field verification of 2–D and 3–D suspended‐sediment models. Journal of Hydraulic Engineering, 116: 12701288.Google Scholar
van Wieren, S. E., and Bakker, J. P. 2008. The impact of browsing and grazing herbivores on biodiversity. In: Gordon, I. J., and Prins, H. H. T., eds. The Ecology of Browsing and Grazing. Springer, Berlin, pp. 236292.Google Scholar
Vandenbruwaene, W., Meire, P., and Temmerman, S. 2012. Formation and evolution of a tidal channel network within a constructed tidal marsh. Geomorphology, 151–152: 114125.Google Scholar
Vu, H. D., and Pennings, S. C. 2018. Predators mediate above- vs. belowground herbivory in a salt marsh crab. Ecosphere, 9(2). doi:10.1002/ecs2.2107.Google Scholar
Vu, H. D., Wieski, K., and Pennings, S. C. 2017. Ecosystem engineers drive creek formation in salt marshes. Ecology, 98: 162174.Google Scholar
Wang, J. Q., Zhang, X. D., Jiang, L. F., Bertness, M. D., Fang, C. M., Chen, J. K., Hara, T., and Li, B. 2010. Bioturbation of burrowing crabs promotes sediment turnover and carbon and nitrogen movements in an estuarine salt marsh. Ecosystems, 13: 586599.Google Scholar
Wang, J. Q., Zhang, Nie, Fu, M., Chen, C. Z., J. K., and Li, B. 2008. Exotic Spartina alterniflora provides compatible habitats for native estuarine crab Sesarma dehaani in the Yangtze River Estuary. Ecological Engineering, 34: 5764.Google Scholar
Wang, M., Gao, X., and Wang, W. 2014. Differences in burrow morphology of crabs between Spartina alterniflora marsh and mangrove habitats. Ecological Engineering, 69: 213219.Google Scholar
Wang, Y. P., Zhang, R., and Gao, S. 1999. Velocity variations in salt marsh creeks, Jiangsu, China. Journal of Coastal Research, 15: 471477.Google Scholar
Watts, C. W., Tolhurst, T. J., Black, K. S., and Whitmore, A. P. 2003. In Situ measurements of erosion shear stress and geotechnical shear strength of the intertidal sediments of the experimental managed realignment scheme at Tollesbury, Essex, UK. Estuarine, Coastal and Shelf Science, 58: 611–20.Google Scholar
Weissburg, M. 1992. Functional analysis of fiddler crab foraging: sex-specific mechanics and constraints in Uca pugnax (Smith). Journal of Experimental Marine Biology and Ecology, 156: 105124.Google Scholar
West, J. M., and Zedler, J. B. 2000. Marsh-creek connectivity: fish use of a tidal salt marsh in Southern California. Estuaries, 23: 699710.CrossRefGoogle Scholar
Widdows, J., and Brinsley, M. 2002. Impact of biotic and abiotic processes on sediment dynamics and the consequences to the structure and functioning of the intertidal zone. Journal of Sea Research, 48: 143156.Google Scholar
Widdows, J., Brinsley, M. D., Bowley, N., and Barrett, C. 1998. A benthic annular flume for in situ measurement of suspension feeding/biodeposition rates and erosion potential of intertidal cohesive sediments. Estuarine Coastal and Shelf Science, 46: 2738.Google Scholar
Widdows, J., Pope, N., and Brinsley, M. 2008. Effect of Spartina anglica stems on nearbed hydrodynamics, sediment erodability and morphological changes on an intertidal mudflat. Marine Ecology Progress Series, 362: 4557.Google Scholar
Wigand, C., Brennan, P., Stolt, M., Holt, M., and Ryba, S. 2009. Soil respiration rates in coastal marshes subject to increasing watershed nitrogen loads in southern New England, US. Wetlands, 29: 952963.CrossRefGoogle Scholar
Wilson, K., Kelley, J., Croitoru, A., Dionne, M., Belknap, D., and Steneck, R. 2009. Stratigraphic and ecophysical characterizations of salt pools: dynamic landforms of the Webhannet Salt Marsh, Wells, ME, USA. Estuaries and Coasts, 32: 855870.Google Scholar
Wilson, K., Kelley, J. T., Tanner, B. R., and Belknap, D. F. 2010. Probing the origins and stratigraphic signature of salt pools from north-temperate marshes in Maine, U.S.A. Journal of Coastal Research, 26: 10071026.Google Scholar
Wilson, C., and Allison, M. 2008. An equilibrium profile model for retreating marsh shorelines in southeast Louisiana. Estuarine, Coastal and Shelf Science, 80, 483494.Google Scholar
Wilson, C. A., Hughes, Z. J., and FitzGerald, D. M. 2012. The effects of crab bioturbation on mid-Atlantic saltmarsh tidal creek extension: geotechnical and geochemical changes. Estuarine, Coastal and Shelf Science, 106: 3344.Google Scholar
Wilson, C. A., Hughes, Z. J., FitzGerald, D. M., Hopkinson, C. S., Valentine, V., and Kolker, A. S. 2014. Saltmarsh pool and tidal creek morphodynamics: dynamic equilibrium of northern latitude saltmarshes? Geomorphology, 213: 99115.Google Scholar
Windham, L. 2001. Comparison of biomass production and decomposition between Phragmites australis (Common Reed) and Spartina patens (Salt Hay Grass) in brackish tidal marshes of New Jersey, USA. Wetlands, 21: 179188.Google Scholar
Xin, P., Jin, G., Li, L. and Barry, D. A. 2009. Effects of crab burrows on pore water flows in salt marshes. Advances in Water Resources, 32: 439449.Google Scholar
Yallop, M. L., Paterson, D. M., and Wellsbury, P. 2000. Interrelationships between rates of microbial production, exopolymer production, microbial biomass, and sediment stability in biofilms of intertidal sediments. Microbial Ecology, 39: 116127.Google Scholar
Yang, S. L., Li, M., Dai, S. B., Liu, Z., Zhang, J. and Ding, P. X. 2006. Drastic decrease in sediment supply from the Yangtze River and its challenge to coastal wetland management. Geophysical Research Letters, 33: 47.Google Scholar
Yang, S. L., Li, H., Ysebaert, T., Bouma, T. J., Zhang, W. X., Wang, Y. Y., Li, P., Li, M., and Ding, P. X. 2008. Spatial and temporal variations in sediment grain size in tidal wetlands, Yangtze Delta: On the role of physical and biotic controls. Estuarine Coastal and Shelf Science, 77: 657671.Google Scholar
Yapp, R. H., Johns, D., and Jones, O. T. 1917. The salt marshes of the Dovey Estuary. Journal of Ecology, 5: 65103.Google Scholar

References

Allen, J. R. L. 1990. Constraints on measurements of sea-level movements from salt-marsh accretion rates. Journal of the Geological Society of London, 147: 57.Google Scholar
Allen, J. R. L. 2000. Morphodynamics of Holocene salt marshes: a review sketch from the Atlantic and Southern North Sea coasts of Europe. Quaternary Science Reviews, 19: 11551231.Google Scholar
Aretxabaleta, A. L., Ganju, N. K., Butman, B., and Signell, R. P. 2017. Observations and a linear model of water level in an interconnected inlet-bay system. Journal of Geophysical Research Oceans, 122: 27602780.Google Scholar
Atwater, B. F., Nelson, A. R., Clague, J. J., Carver, G. A., Bobrowsky, P. T., Bourgeois, J., and Darienzo, M. E. 1995. Summary of coastal geologic evidence for past great earthquakes at the Cascadia subduction zone. Earthquake Spectra, 11: 118.Google Scholar
Avnaim-Katav, S., Gehrels, W. R., Brown, L. N., Fard, E., and MacDonald, G. M. 2017. Distributions of salt-marsh foraminifera along the coast of SW California, USA: implications for sea-level reconstructions. Marine Micropalaeontology, 131: 2543.Google Scholar
Barlow, N. L. M., Long, A. J., Saher, M. H., Gehrels, W. R., Garnett, M. H., and Scaife, R. G. 2014. Salt-marsh reconstructions of relative sea-level change in the North Atlantic during the last 2000 years. Quaternary Science Reviews, 99: 116.Google Scholar
Barlow, N. L. M., Shennan, I., Long, A. J., Gehrels, W. R., Saher, M., Woodroffe, S. A., and Hillier, C. 2013. Salt marshes as late Holocene tide gauges. Global and Planetary Change, 106: 90110.CrossRefGoogle Scholar
Barnett, R. L., Garneau, M., and Bernatchez, P. 2016. Salt-marsh sea-level indicators and transfer function development for the Magdalen Islands in the Gulf of St. Lawrence, Canada. Marine Micropaleontology, 122: 1326.Google Scholar
Barnett, R. L., Gehrels, W. R., Charman, D. J., Saher, M. H., and Marshall, W. A. 2015. Late Holocene sea-level change in Arctic Norway. Quaternary Science Reviews, 107: 214230.Google Scholar
Barnett, R. L., Newton, T. L., Charman, D. J., and Gehrels, W. R. 2017. Salt-marsh testate amoebae as precise and widespread indicators of sea-level change. Earth Science Reviews, 164: 193207.Google Scholar
Bartram, W. 1791. Travels through North and South Carolina, Georgia, East and West Florida. Philadelphia, James and Johnson.Google Scholar
Bloom, A. L., and Stuiver, M. 1963. Submergence of the Connecticut coast. Science, 139: 332334.Google Scholar
Bradley, W. H. 1953. Age of intertidal tree stumps at Robinhood, Maine. American Journal of Science, 251: 543546.Google Scholar
Brain, M. J., Kemp, A. C., Hawkes, A. D., Engelhart, S. E., Vane, C. H., Cahill, N., Hill, T. D., Donnelly, J. P., and Horton, B. P. 2017. Exploring mechanisms of compaction in salt-marsh sediments using Common Era relative sea-level reconstructions. Quaternary Science Reviews, 167: 96111.Google Scholar
Brain, M. J., Kemp, A. C., Horton, B. P., Culver, S. J., Parnell, A. C., and Cahill, N. 2015. Quantifying the contribution of sediment compaction to late Holocene salt-marsh sea-level reconstructions, North Carolina, USA. Quaternary Research, 83: 4151.Google Scholar
Cahill, N., Kemp, A. C., Horton, B. P., and Parnell, A. C. 2016. A Bayesian hierarchical model for reconstructing relative sea level: from raw data to rates of change. Climate of the Past, 12: 525542.Google Scholar
Callard, S. L., Gehrels, W. R., Morrison, B. V. and Grenfell, H. R. 2011. Suitability of salt-marsh foraminifera as proxy indicators of sea level in Tasmania. Marine Micropaleontology, 79: 121131.Google Scholar
Chapman, V. J. 1940. Succession on the New England salt marshes. Ecology, 21: 279282.Google Scholar
Charman, D. J., Roe, H. M., and Gehrels, W. R. 2002. Modern distribution of saltmarsh testate amoebae: regional variability of zonation and response to environmental variables. Journal of Quaternary Science, 17: 387409.Google Scholar
Donnelly, J. P., Cleary, P., Newby, P., and Ettinger, R. 2004. Coupling instrumental and geological records of sea-level change: evidence from southern New England of an increase in the rate of sea-level rise in the 19th century. Geophysical Research Letters, 30, doi:10.1029/2003GL017801.Google Scholar
Edwards, R., and Wright, A. 2015. Foraminifera. In: Shennan, I., Long, A. J., and Horton, B. P., eds., Handbook of Sea-Level Research, Wiley, Chichester, pp. 191217.Google Scholar
Edwards, R. J., van de Plassche, O., Gehrels, W. R., and Wright, A. J. 2004. Assessing sea-level data from Connecticut, USA, using a foraminiferal transfer function for tide level. Marine Micropalaeontology, 51: 239255.Google Scholar
Eleuterius, L. N. 1976. The distribution of Juncus roemerianus in the salt marshes of North America. Chesapeake Science, 17: 289292.Google Scholar
Engelhart, S. E., Horton, B. P., and Kemp, A. C. 2011b. Holocene sea-level changes along the United States’ Atlantic Coast. Oceanography, 24: 7079.Google Scholar
Engelhart, S. E., Horton, B. P., Nelson, A. R., Hawkes, A. D., Witter, R. C., Wang, K., Wang, P.-L., and Vane, C. H. 2013. Testing the use of microfossils to reconstruct great earthquakes at Cascadia. Geology, 41: 10671070.Google Scholar
Engelhart, S. E., Peltier, W. R., and Horton, B. P. 2011a. Holocene relative sea- level changes and glacial isostatic adjustment of the U.S. Atlantic coast. Geology, 39: 751754.Google Scholar
Engelhart, S. E., Vacchi, M., Horton, B. P., Nelson, A. R., and Kopp, R. E. 2015. A sea-level database for the Pacific coast of central North America. Quaternary Science Reviews, 113: 7892.Google Scholar
Fatela, F., Taborda, R. 2002. Confidence limits of species proportions in microfossil assemblages. Marine Micropaleontology, 45: 169174.Google Scholar
French, J. R., Spencer, T., Murray, A. L., and Arnold, N. S. 1995. Geostatistical analysis of sediment deposition in two small tidal wetlands. Journal of Coastal Research, 11: 308321.Google Scholar
Gehrels, W. R. 1994. Determining relative sea-level change from salt-marsh foraminifera and plant zones on the coast of Maine, U.S.A. Journal of Coastal Research, 10: 9901009.Google Scholar
Gehrels, W. R. 1999. Middle and late Holocene sea-level changes in eastern Maine reconstructed from foraminiferal saltmarsh stratigraphy and AMS 14C dates on basal peat. Quaternary Research, 52: 350359.Google Scholar
Gehrels, W. R. 2000. Using foraminiferal transfer functions to produce high-resolution sea-level records from saltmarsh deposits, Maine, USA. The Holocene, 10: 367376.Google Scholar
Gehrels, W. R. 2002. Intertidal foraminifera as palaeoenvironmental indicators. In: Haslett, S. K., ed., Quaternary Environmental Micropalaeontology, Arnold Publishers, New York, pp. 91114.Google Scholar
Gehrels, W. R., and Belknap, D. F. 1993. Neotectonic history of eastern Maine evaluated from historic sea-level data and 14C dates on salt-marsh peats. Geology, 21: 615618.Google Scholar
Gehrels, W. R., Belknap, D. F., and Kelley, J. T. 1996. Integrated high-precision analyses of Holocene relative sea-level changes: Lessons from the coast of Maine. Geological Society of America Bulletin, 108: 10731088.Google Scholar
Gehrels, W. R., Callard, S. L., Moss, P. T., Marshall, W. A., Blaauw, M., Hunter, J., Milton, J. A., and Garnett, M. H. 2012. Nineteenth and twentieth century sea-level changes in Tasmania and New Zealand. Earth and Planetary Science Letters, 315–316: 94102.Google Scholar
Gehrels, W. R., Kirby, J. R., Prokoph, A., Newnham, R. M., Achterberg, E. P., Evans, E. H., Black, S., and Scott, D. B. 2005. Onset of recent rapid sea-level rise in the western Atlantic Ocean. Quaternary Science Reviews, 24: 20832100.CrossRefGoogle Scholar
Gehrels, W. R., Marshall, W. A., Gehrels, M. J., Larsen, G., Kirby, J. R., Eiriksson, J., Heinemeier, J., and Shimmield, T. 2006. Rapid sea-level rise in the North Atlantic Ocean since the first half of the 19th century. The Holocene, 16: 948964. Erratum, The Holocene 17: 419-420.Google Scholar
Gehrels, W. R., Roe, H. M., and Charman, D. J. 2001. Foraminifera, testate amoebae and diatoms as sea-level indicators in UK saltmarshes: a quantitative multiproxy approach. Journal of Quaternary Science, 16: 201220.Google Scholar
Gehrels, W. R., and Shennan, I. 2015. Sea level in time and space: revolutions and inconvenient truths. Journal of Quaternary Science, 30: 131143.Google Scholar
Gerlach, M. J., Engelhart, S. E., Kemp, A. C., Moyer, R. P., Smoak, J. M., Bernhardt, C. E., and Cahill, N. 2017. Reconstructing Common Era relative sea-level change on the Gulf Coast of Florida. Marine Geology, 390: 254269.Google Scholar
González, J. L. and Törnqvist, T. E. 2009. A new Late Holocene sea-level record from the Mississippi Delta: Evidence for a climate/sea level connection? Quaternary Science Reviews, 28, 17371749.Google Scholar
Goslin, J., Sansjofre, P., Van Vliet-Lanoë, B., and Delacourt, C. 2017. Carbon stable isotope δ13C) and elemental TOC, TN) geochemistry in saltmarsh surface sediments Western Brittany, France): a useful tool for reconstructing Holocene relative sea-level. Journal of Quaternary Science, 32: 9891007.Google Scholar
Guilbault, J. -P., Clague, J. J., and Lapointe, M. 1995. Amount of subsidence during a late Holocene earthquake – evidence from fossil tidal marsh foraminifera at Vancouver Island, west coast of Canada. Palaeogeography, Palaeoclimatology, Palaeoecology, 118: 4971.Google Scholar
Hawkes, A. D., Horton, B. P., Nelson, A. R., and Hill, D. F. 2010. The application of intertidal foraminifera to reconstruct coastal subsidence during the giant Cascadia earthquake of AD 1700 in Oregon, USA. Quaternary International, 221: 116140.Google Scholar
Hawkes, A. D., Horton, B. P., Nelson, A. R., Vane, C. H., and Sawai, Y. 2011. Coastal subsidence in Oregon, USA, during the giant Cascadia earthquake of AD 1700. Quaternary Science Reviews, 30: 364376.Google Scholar
Horton, B. P. 1999. The distribution of contemporary intertidal foraminifera at Cowpen Marsh, Tees Estuary,UK: implications for studies of Holocene sea-level changes. Palaeogeography, Palaeoclimatology, Palaeoecology, 149: 127149.Google Scholar
Horton, B. P., and Edwards, R. J. 2006. Quantifying Holocene sea level change using intertidal foraminifera: lessons from the British Isles. Retrieved from http://repository.upenn.edu/ees_papers/50.CrossRefGoogle Scholar
Horton, B. P., Edwards, R. J., and Lloyd, J. M. 1999. A foraminiferal-based transfer function: Implications for sea- level studies. Journal of Foraminiferal Research, 29: 117129.Google Scholar
Horton, B. P., Milker, Y., Dura, T., Wang, K., Bridgeland, W. T., Brophy, L., Ewald, M., et al. 2017. Microfossil measures of rapid sea-level rise: timing of response of two microfossil groups to a sudden tidal-flooding experiment in Cascadia. Geology, 45: 535538.Google Scholar
Imbrie, J., and Kipp, N. G. 1971. A new micropaleontological method for quantitative paleoclimatology: Application to a late Pleistocene Caribbean core. In: Turekian, K. K., ed., The Late Cenozoic Glacial Ages, Yale University Press, New Haven, pp. 71181.Google Scholar
Jelgersma, S. 1961. Holocene sea-level changes in the Netherlands. Mededelingen Geologische Stichting C-IV, 7: 1100.Google Scholar
Kaye, C. A., and Barghoorn, E. S. 1964. Late Quaternary sea-level change and crustal rise at Boston, Massachusetts, with notes on the autocompaction of peat. Geological Society of America Bulletin, 75: 6380.Google Scholar
Kemp, A. C., Cahill, N., Engelhart, S. E., Hawkes, A. E., and Wang, K. 2018. Revising estimates of spatially variable subsidence during the A.D. 1700 Cascadia earthquake using a Bayesian foraminiferal transfer function. Bulletin of the Seismological Society of America, 108: 654673.Google Scholar
Kemp, A. C., Engelhart, S. E., Culver, S. J., Nelson, A., Briggs, R. W., and Haeussler, P. J. 2013. Modern salt-marsh and tidal-flat foraminifera from Sitkinak and Simeonof Islands, southwestern Alaska. Journal of Foraminiferal Research, 43: 8898.Google Scholar
Kemp, A. C., Horton, B. P., Culver, S. J., Corbett, D. R., van de Plassche, O., Gehrels, W. R., Douglas, B. C., and Parnell, A. C. 2009. Timing and magnitude of recent accelerated sea-level rise (North Carolina, United States). Geology, 37: 10351038.Google Scholar
Kemp, A. C., Horton, B. P., Donnelly, J. P., Mann, M. E., Vermeer, M., and Rahmstorf, S. 2011. Climate related sea-level variations over the past two millennia. Proceedings of the National Academy of Sciences of the USA, 108: 1101711022.Google Scholar
Kemp, A. C., Horton, B. P., Nikitina, D., Vane, C. H., Potapova, M., Weber-Bruya, E., Culver, S. J., Repkina, T., and Hill, D. F. 2017a. The distribution and utility of sea-level indicators in Eurasian sub-Arctic salt marshes White Sea, Russia Boreas, 46: 562584.Google Scholar
Kemp, A. C., Kegel, J. J., Culver, S. J., Barber, D. C., Mallinson, D. J., Leorri, E., Bernhardt, C. E., Cahill, N., et al. 2017b. Extended late Holocene relative sea-level histories for North Carolina, USA. Quaternary Science Reviews, 160: 1330.Google Scholar
Kemp, A. C., Sommerfield, C. K., Vane, C. H., Horton, B. P., Chenery, S., Anisfeld, S., and Nikitina, D. 2012. Use of lead isotopes for developing chronologies in recent salt-marsh sediments. Quaternary Geochronology, 12: 4049.Google Scholar
Kemp, A. C., Telford, R. J. 2015. Transfer functions. In: Shennan, I., Long, A. J., and Horton, B. P., eds. Handbook of Sea-Level Research, Wiley, Chichester, pp. 470499.Google Scholar
Kemp, A. C., Vane, C. H., Horton, B. P., Engelhart, S. E., and Nikitina, D. 2012. Application of stable carbon isotopes for reconstructing salt-marsh floral zones and relative sea level, New Jersey, USA. Journal of Quaternary Science, 27: 404414.Google Scholar
Kemp, A. C., Wright, A. J., Barnett, R. L., Hawkes, A. D., Charman, D. J., Sameshima, C., King, A. N., et al. 2017c. Utilty of salt-marsh foraminifera, testate amoebae and bulk-sediment δ13C values as sea-level indicators in Newfoundland, Canada. Marine Micropaleontology, 130: 4359.Google Scholar
Kemp, A. C., Wright, A. J., Edwards, R. J., Barnett, R. L., Brain, M. J., Kopp, R. E., Cahill, N., et al. 2018. Relative sea-level change in Newfoundland, Canada during the past ~3000 years. Quaternary Science Reviews, 201: 89110.Google Scholar
Kirwan, M. L., Temmerman, S., Skeehan, E. E., Guntenspergen, G. R., and Fagherazzi, S. 2016. Overestimation of marsh vulnerability to sea level rise. Nature Climate Change, 6: 253260.Google Scholar
Kopp, R. E., Kemp, A. C., Bittermann, K., Horton, B. P., Donnelly, J. P., Gehrels, W. R., Hay, , et al. 2016. Temperature-driven global sea-level variability in the Common Era. Proceedings of the Natural Academy of Sciences of the United States of America, 113: E1434E1441.Google Scholar
Lamb, A. L., Wilson, G. P., and Leng, M. L. 2006. A review of coastal palaeoclimate and relative sea-level reconstructions using δ13C and C/N ratios in organic material. Earth-Science Reviews, 75: 2957.CrossRefGoogle Scholar
Le Coze, F., and Hayward, B. 2017. Entzia macrescens Brady, 1870 In: Hayward, B. W., Le Coze, F., and Gross, O. World Foraminifera Database. Accessed at www.marinespecies.org/foraminifera/aphia.php?p=taxdetails&id=742429 on 2017-10-26.Google Scholar
Libby, W. F. 1961. Radiocarbon dating. Science, 133: 621629.Google Scholar
Long, A. J., Barlow, N. L. M., Gehrels, W. R., Saher, M. H., Woodworth, P. L., Scaife, R. G., Brain, M. J., and Cahill, N. 2014. Contrasting records of sea-level change in the eastern and western North Atlantic during the last 300 years. Earth and Planetary Science Letters, 388: 110122.Google Scholar
Lyell, C. 1849. A Second Visit to the United States of North America, in Two Volumes. Volume 1. New York, Harper and Brothers.Google Scholar
Marshall, W. A., Gehrels, W. R., Garnett, M. H., Freeman, S. P. H. T., Maden, C., and Xu, S. 2007. The use of “bomb spike” calibration and high-precision AMS 14C analyses to date salt-marsh sediments deposited during the past three centuries. Quaternary Research, 68: 325337.Google Scholar
Mudge, B. F. 1858. The salt marsh formations of Lynn. Proceedings of the Essex Institute, 2: 117119.Google Scholar
Murray, J. W. 1982. Benthic foraminifera: the variability of living, dead or total assemblages in the interpretation of palaeoecology. Journal of Micropalaeontology, 1: 137140.Google Scholar
Nelson, A. R., Shennan, I., and Long, A. J. 1995. Identifying coseismic subsidence in tidal-wetland stratigraphic sequences at the Cascadia subduction zone of western North America. Journal of Geophysical Research, 101: 61156135.Google Scholar
Nikitina, D. L., Kemp, A. C., Horton, B. P., Vane, C. H., van de Plassche, O., and Engelhart, S. E. 2014. Storm erosion during the past 2000 years along the north shore of Delaware Bay, USA. Geomorphology, 208: 160172.Google Scholar
Parnell, A. C., and Gehrels, W. R. 2015. Using chronological models in late Holocene sea level reconstructions from saltmarsh sediments. In: Shennan, I., Long, A. J., and Horton, B. P., eds. Handbook of Sea-Level Research, Wiley, Chichester, pp. 500513.Google Scholar
Patterson, R. T., and Fishbein, E. 1989. Re-examination of the statistical methods used to determine the number of point counts needed for micropaleontological quantitative research. Journal of Palaeontology, 63: 245248.Google Scholar
Payne, R. J., and Mitchell, E. A. D. 2009. How many is enough? Determining optimal count totals for ecological and palaeoecological studies of testate amoebae. Journal of Paleolimnology, 42: 483495.Google Scholar
Piecuch, C. G., Bittermann, K., Kemp, A. C., Ponte, R. M., Little, C. M., Engelhart, S. E., and Lentz, S. J. 2018. River-discharge effects on United States Atlantic and Gulf coast sea-level changes. Proceedings of the National Academy of Sciences of the USA, 30: 77297734.Google Scholar
Redfield, A. C. 1959. The Barnstable marsh. In Ragotzkie, R. A., Pomeroy, L. R., Teal, J. M., and Scott, D. C., eds., Proceedings, Salt marsh Conference, May 25–28, 1958, Sapelo Island, Athens, Georgia, University of Georgia, pp. 3742.Google Scholar
Redfield, A. C. 1972. Development of a New England salt marsh. Ecological Monographs, 42: 210237.Google Scholar
Redfield, A. C., and Rubin, M. 1962. The age of salt marsh peat and its relations to recent change in sea level at Barnstable, Massachusetts. Proceedings of the National Academy of Sciences of the USA, 48: 17281735.Google Scholar
Reimer, P. J., Bard, E., Bayliss, A., Beck, J. W., Blackwell, P. G., Bronk Ramsey, C., Buck, C. E., et al. 2013. IntCal13 and Marine13 radiocarbon age calibration curves 0–50,000 years cal BP. Radiocarbon, 55: 18691887.Google Scholar
Roe, H. M., Charman, D. J., and Gehrels, W. R. 2002. Fossil testate amoebae in coastal deposits in the UK: implications for studies of sea-level change. Journal of Quaternary Science, 17: 411429.Google Scholar
Sachs, H. M. 1977. Paleoecological transfer functions. Annual Review of Earth and Planetary Sciences, 5: 159178.Google Scholar
Saher, M. H., Gehrels, W. R., Barlow, N. L. M., Long, A. J., Haigh, I. D., and Blaauw, M. 2015. A 600-year multiproxy record of sea-level change and the influence of the North Atlantic Oscillation. Quaternary Science Reviews, 108: 2336.Google Scholar
Scott, D. B., and Medioli, F. S. 1978. Vertical zonations of marsh foraminifera as accurate indicators of former sea-levels. Nature, 272: 528531.Google Scholar
Scott, D. B., and Medioli, F. S. 1980a. Quantitative studies of marsh foraminifera distribution in Nova Scotia: implications for sea-level studies. Cushman Foundation for Foraminiferal Research Special Publication, 17: 158.Google Scholar
Shaler, N. S. 1886. Preliminary report on sea-coast swamps of the eastern United States. US Geological Survey 6th Annual Report, pp. 353–398.Google Scholar
Shennan, I., and Horton, B. 2002. Holocene land- and sea-level changes in Great Britain. Journal of Quaternary Science, 17: 511526.Google Scholar
Shennan, I., Long, A.J., and Horton, B.P., eds. 2015. Handbook of Sea-Level Research, Wiley, Chichester.Google Scholar
Streif, H. 1979. Cyclic formation of coastal deposits and their indications of vertical sea-level changes. Oceanus, 5: 303306.Google Scholar
Strachan, K. L., Hill, T. R., Finch, J. M., and Barnett, R. L. 2015. Vertical zonation of foraminifera assemblages in Galpins Salt Marsh, South Africa. Journal of Foraminiferal Research, 45: 2941.Google Scholar
Szkornik, K., Gehrels, W. R., and Murray, A. S. 2008. Aeolian sand movement and relative sea-level rise in Ho Bugt, western Denmark, during the Little Ice Age. The Holocene, 18: 951965.Google Scholar
Törnqvist, T. E., van Ree, M. H. M., van 't Veer, R., and van Geel, B. 1998. Improving methodology for high-resolution reconstruction of sea-level rise and neotectonics by paleoecological analysis and AMS 14C dating of basal peats. Quaternary Research, 49: 7285.Google Scholar
Vacchi, M., Engelhart, S. E., Nikitina, D., Ashe, E. L., Peltier, W. R., Roy, K., Kopp, R. E., and Horton, B. P. 2018. Postglacial relative sea-level histories along the eastern Canadian coastline. Quaternary Science Reviews, 201: 124146.Google Scholar
Vacchi, M., Marriner, N., Morhange, C., Spada, G., Fontana, A., and Rovere, A. 2016. Multiproxy assessment of Holocene relative sea-level changes in the western Mediterranean: sea-level variability and improvements in the definition of the isostatic signal. Earth-Science Reviews, 155: 172197.Google Scholar
Van de Plassche, O. 2000. North Atlantic climate-ocean variations and sea level in Long Island Sounds, Connecticut, since 500 cal yr AD. Quaternary Research, 53: 8997.Google Scholar
Van de Plassche, O., Mook, W. G., and Bloom, A. L. 1989. Submergence of coastal Connecticut 6000-3000 14C) years B.P. Marine Geology, 86: 349354.Google Scholar
Van de Plassche, O., Wright, A. J., van der Borg, K., and de Jong, A. F. M. 2004. On the erosive trail of a 14th and 15th century hurricane in Connecticut USA) salt marshes. Radiocarbon, 46: 775784.Google Scholar
Van der Wal, D., and Pye, K. 2002. Patterns, rates and possible causes of saltmarsh erosion in the Greater Thames area UK. Geomorphology, 61: 373391.Google Scholar
Watcham, E. P., Shennan, I., and Barlow, N. L. M. 2013. Scale considerations in using diatoms as indicators of sea-level change: lessons from Alaska. Journal of Quaternary Science, 28: 165179.Google Scholar
Wells, B. W. 1928. Plant communities of the coastal plain of North Carolina and their successional relations. Ecology, 9: 230242.Google Scholar
Wilson, G. P. 2017. On the application of contemporary bulk sediment organic carbon isotope and geochemical datasets for Holocene sea-level reconstruction in NW Europe. Geochimica et Cosmochimica Acta, 214: 191208.Google Scholar
Woodroffe, S. A. 2009. Testing models of mid to late Holocene sea-level change, North Queensland, Australia. Quaternary Science Reviews, 28: 24742488.Google Scholar
Wright, A. J., Edwards, R. J., and van de Plassche, O. 2011. Reassessing transfer-function performance in sea-level reconstruction based on benthic salt-marsh foraminifera from the Atlantic coast of NE North America. Marine Micropaleontology, 81: 4362.Google Scholar
Zoccorato, C., and Teatini, P. 2017. Numerical simulations of Holocene salt-marsh dynamics under the hypothesis of large soil deformations. Advances in Water Resources, 110: 107119.Google Scholar

References

Allen, J. R. L. 1989. Evolution of salt‐marsh cliffs in muddy and sandy systems: A qualitative comparison of British West‐Coast estuaries. Earth Surface Processes and Landforms, 14: 8592.Google Scholar
Allison, M. A., and Kepple, E. 2001. Modern sediment supply to the lower delta plain of the Ganges-Brahmaputra River in Bangladesh. Geo-Marine Letters, 21: 6674.Google Scholar
Barbier, E. B., Georgiou, I. Y., Enchelmeyer, B., and Reed, D. J. 2013. The value of wetlands in protecting Southeast Louisiana from hurricane storm surges. PLOS ONE, 8: 16.Google Scholar
Barbier, E. B., Hacker, S. D., Kennedy, C., Koch, E. W., Stier, A. C., and Silliman, B. R. 2011. The value of estuarine and coastal ecosystem services. Ecological Monographs, 81: 169193.Google Scholar
Barras, J. A. 2007. Land area changes in coastal Louisiana after Hurricanes Katrina and Rita. In Farris, G. S., Smith, G. J., Crane, M. P., Demas, C. R., Robbins, L. L., and Lavoie, D. L., eds., Science and the Storms: The USGS Response to the Hurricanes of 2005 pp. 97–112. U.S. Geological Survey Circular 1306.Google Scholar
Baustian, J. J., and Mendelssohn, I. A. 2015. Hurricane-induced sedimentation improves marsh resilience and vegetation vigor under high rates of relative sea level rise. Wetlands, 35: 795802.Google Scholar
BenDor, T., Lester, T. W., Livengood, A., Davis, A., and Yonavjak, L. 2015. Estimating the size and impact of the ecological restoration economy. PLOS ONE, 10: 115. https://doi.org/10.1371/journal.pone.0128339Google Scholar
Boldt, K. V., Lane, P., Woodruff, J. D., and Donnelly, J. P. 2010. Calibrating a sedimentary record of overwash from Southeastern New England using modeled historic hurricane surges. Marine Geology, 275: 127139.Google Scholar
Boose, E. R., Chamberlin, K. E., and Foster, D. R. 2001. Landscape and regional impacts of hurricanes in New England. Ecological Monographs, 71: 2748.Google Scholar
Brandon, C. M., Woodruff, J. D., Lane, D. P., and Donnelly, J. P. 2013. Tropical cyclone wind speed constraints from resultant storm surge deposition: A 2500 year reconstruction of hurricane activity from St. Marks, FL. Geochemistry, Geophysics, Geosystems, 14: 29933008.Google Scholar
Burkett, V., Groat, C. G., and Reed, D. 2007. Hurricanes not the key to a sustainable coast. Science, 315: 13661367.Google Scholar
Cahoon, D. R. 2006. A review of major storm impacts on coastal wetland elevations. Estuaries and Coasts, 29: 889898.Google Scholar
Costanza, R., d’Arge, R., de Groot, R., Farber, S., Grasso, M., Hannon, B., van den Belt, M.. Limburg, K. et al. 1997. The value of the world’s ecosystem services and natural capital. Nature, 387: 253260.Google Scholar
Costanza, R., de Groot, R., Sutton, P., van der Ploeg, S., Anderson, S. J., Kubiszewski, I., Farber, S., and Turner, R. K. 2014. Changes in the global value of ecosystem services. Global Environmental Change, 26: 152158.Google Scholar
Craft, C., Clough, J., Ehman, J., Jove, S., Park, R., Pennings, S., Guo, H. and Machmuller, M. 2009. Forecasting the effects of accelerated sea-level rise on tidal marsh ecosystem services. Frontiers in Ecology and the Environment, 7: 7378.Google Scholar
Day, J. W., Boesch, D. F., Clairain, E. J., Kemp, G. P., Laska, S. B., Mitsch, W. J., Orth, K. et al. 2007. Restoration of the Mississippi Delta: Lessons from Hurricanes Katrina and Rita. Science, 315: 16791684.Google Scholar
Donnelly, J. P. 2004. Coupling instrumental and geological records of sea-level change: Evidence from southern New England of an increase in the rate of sea-level rise in the late 19th century. Geophysical Research Letters, 31: 25.Google Scholar
Donnelly, J. P., Hawkes, A. D., Lane, P., Macdonald, D., Shuman, B. N., Toomey, M. R., van Hengstum, P. J., and Woodruff, J. D. 2015. Climate forcing of unprecedented intense-hurricane activity in the last 2000 years. Earth’s Future, 3: 4965.Google Scholar
Donnelly, J. P., Roll, S., Wengren, M., Butler, J., Lederer, R., and Webb, T. 2001. Sedimentary evidence of intense hurricane strikes from New Jersey. Geology, 29: 615618.Google Scholar
Donnelly, J. P., Smith Bryant, S., Butler, J., Dowling, J., Fan, L., Hausmann, N., Newby, P., et al. 2001. 700 yr sedimentary record of intense hurricane landfalls in southern New England. Geological Society of America Bulletin, 113: 714727.Google Scholar
Elsey-Quirk, T. 2016. Impact of Hurricane Sandy on salt marshes of New Jersey. Estuarine, Coastal and Shelf Science, 183: 235248.Google Scholar
Engelhart, S. E., Horton, B. P., Douglas, B. C., Peltier, W. R., and Törnqvist, T. E. 2009. Spatial variability of late Holocene and 20th century sea-level rise along the Atlantic coast of the United States. Geology, 37: 11151118.Google Scholar
Fagherazzi, S., Mariotti, G., Wiberg, P. L., and McGlathery, K. J. 2013. Marsh collapse does not require sea level rise. Oceanography, 26: 7077.Google Scholar
Feagin, R. A., Lozada-Bernard, S. M., Ravens, T. M., Moller, I., Yeager, K. M., and Baird, A. H. 2009. Does vegetation prevent wave erosion of salt marsh edges? Proceedings of the National Academy of Sciences of the USA, 106: 1010910113.Google Scholar
French, J. 2006. Tidal marsh sedimentation and resilience to environmental change: Exploratory modelling of tidal, sea-level and sediment supply forcing in predominantly allochthonous systems. Marine Geology, 235: 119136.Google Scholar
Frey, R. W., and Basan, P. B. 1978. Coastal salt marshes. In Davis, R. A. Jr., ed., Coastal Sedimentary Environments. New York: Springer, pp. 225302.Google Scholar
Ganju, N. K., Kirwan, M. L., Dickhudt, P. J., Guntenspergen, G. R., Cahoon, D. R., and Kroeger, K. D. 2015. Sediment transport-based metrics of wetland stability. Geophysical Research Letters, 42: 79928000.Google Scholar
Gardner, L. R., Michener, W. K., Kjerve, B., and Lipscomb, D. J. 1992. Disturbance effects of Hurricane Hugo on a pristine coastal landscape: North Inlet, South Carolina, USA. Netherlands Journal of Sea Reasearch, 30: 249263.Google Scholar
Gedan, K. B., Altieri, A. H., and Bertness, M. D. 2011. Uncertain future of New England salt marshes. Marine Ecology Progress Series, 434: 229237.Google Scholar
Goodbred, S. L., and Hine, A. C. 1995. Coastal storm deposition: Salt-marsh response to a severe extratropical storm, March 1993, west-central Florida. Geology, 23: 679682.Google Scholar
Hippensteel, S. P. 2008. Preservation potential of storm deposits in South Carolina back-barrier marshes. Journal of Coastal Research, 243: 594601.Google Scholar
Hippensteel, S. P., and Martin, R. E. 1999. Foraminifera as an indicator of overwash deposits, Barrier Island sediment supply, and Barrier Island evolution: Folly Island, South Carolina. Palaeogeography, Palaeoclimatology, Palaeoecology, 149: 115125.Google Scholar
Howes, N. C., FitzGerald, D. M., Hughes, Z. J., Georgiou, I. Y., Kulp, M. A., Miner, M. D., Smith, J. M., Barras, J. A. 2010. Hurricane-induced failure of low salinity wetlands. Proceedings of the National Academy of Sciences of the USA, 107: 1401414019.Google Scholar
Hu, K., Chen, Q., Wang, H., Hartig, E. K., and Orton, P. M. 2018. Numerical modeling of salt marsh morphological change induced by Hurricane Sandy. Coastal Engineering, 132: 6381.Google Scholar
Kemp, A. C., Bernhardt, C. E., Horton, B. P., Kopp, R. E., Vane, C. H., Peltier, W. R., Hawkes, A. D., et al. 2014. Late Holocene sea- and land-level change on the U.S. southeastern Atlantic coast. Marine Geology, 357: 90100.Google Scholar
Kemp, A. C., Hawkes, A. D., Donnelly, J. P., Vane, C. H., Horton, B. P., Hill, T. D., Anisfeld, S. C. et al. 2015. Relative sea-level change in Connecticut USA ) during the last 2200 yrs. Earth and Planetary Science Letters, 428: 217229.Google Scholar
Kiage, L., Deocampo, D., Mccloskey, T. A., Bianchette, T. A., and Hursey, M. 2011. A 1900-year paleohurricane record from Wassaw Island, Georgia, USA. Journal of Quaternary Science, 26: 714722.Google Scholar
Kirwan, M. L., Temmerman, S., Skeehan, E. E., Guntenspergen, G. R., and Faghe, S. 2016. Overestimation of marsh vulnerability to sea level rise. Nature Climate Change, 6: 253260.Google Scholar
Kolker, A. S., Goodbred, S. L., Hameed, S., and Cochran, J. K. 2009. High-resolution records of the response of coastal wetland systems to long-term and short-term sea-level variability. Estuarine, Coastal and Shelf Science, 84: 493508.Google Scholar
Kopp, R. E., Kemp, A. C., Bittermann, K., Horton, B. P., Donnelly, J. P., Gehrels, W. R., Hay, C. C., et al. 2016. Temperature-driven global sea-level variability in the Common Era. Proceedings of the National Academy of Sciences of the USA, 113: 18.Google Scholar
van de Koppel, J., van der Wal, D., Bakker, J. P., and Herman, P. M. J. 2005. Self-organization and vegetation collapse in salt marsh ecosystems. The American Naturalist, 165: E1E12.Google Scholar
Lane, P., Donnelly, J. P., Woodruff, J. D., and Hawkes, A. D. 2011. A decadally-resolved paleohurricane record archived in the late Holocene sediments of a Florida sinkhole. Marine Geology, 287: 1430.Google Scholar
Leonard, L. A., Hine, A. C., and Luther, M. E. 1995. Surficial sediment transport and deposition processes in a Juncus roemerianus marsh. Journal of Coastal Research, 11: 322336.Google Scholar
Leonardi, N., Defne, Z., Ganju, N. K., and Fagherazzi, S. 2016. Salt marsh erosion rates and boundary features in a shallow bay. Journal of Geophysical Research: Earth Surface, 121: 18611875.Google Scholar
Leonardi, N., and Fagherazzi, S. 2014. How waves shape salt marshes. Geology, 42: 887890.Google Scholar
Leonardi, N., Ganju, N. K., and Fagherazzi, S. 2016. A linear relationship between wave power and erosion determines salt-marsh resilience to violent storms and hurricanes. Proceedings of the National Academy of Sciences of the USA, 113: 6468.Google Scholar
Loder, N. M., Irish, J. L., Cialone, M. A., and Wamsley, T. V. 2009. Sensitivity of hurricane surge to morphological parameters of coastal wetlands. Estuarine, Coastal and Shelf Science, 84: 625636.Google Scholar
Long, A. J., Waller, M. P., and Stupples, P. 2006. Driving mechanisms of coastal change: Peat compaction and the destruction of late Holocene coastal wetlands. Marine Geology, 2251–4: 6384.Google Scholar
Ludlum, D. M. 1963. Early American Hurricanes. Boston, MA: American Meterological Society.Google Scholar
Marsooli, R., Orton, P. M., Georgas, N., and Blumberg, A. F. 2016. Three-dimensional hydrodynamic modeling of coastal flood mitigation by wetlands. Coastal Engineering, 111: 8394.Google Scholar
McowenC. J.WeatherdonL. V.BochoveJ.-W. V.  SullivanE., BlythS.ZocklerC.Stanwell-Smith, D., et al. 2017. A global map of saltmarshes. Biodiversity Data Journal510.3897/BDJ.5.e11764.Google Scholar
McKee, K. L., and Cherry, J. A. 2009. Hurricane Katrina sediment slowed elevation loss in subsiding brackish marshes of the Mississippi River delta. Wetlands, 29: 215.Google Scholar
McLoughlin, S. M., Wiberg, P. L., Safak, I., and McGlathery, K. J. 2015. Rates and forcing of marsh edge erosion in a shallow coastal bay. Estuaries and Coasts, 38: 620638.Google Scholar
Mendelsohn, R., Emanuel, K., Chonabayashi, S., and Bakkensen, L. 2012. The impact of climate change on global tropical cyclone damage. Nature Climate Change, 2: 205209.Google Scholar
Millennium Ecosystem Assessment 2005. Ecosystems and human well-being: Wetlands and water synthesis. Millennium Ecosystem Assessment. Washington, DC.Google Scholar
Miller, K. G., Sugarman, P. J., Browning, J. V., Horton, B. P., Stanley, A., Kahn, A., Uptegrove, J., and Aucott, M. 2009. Sea-level rise in New Jersey over the past 5000 years: Implications to anthropogenic changes. Global and Planetary Change, 66: 1018.Google Scholar
Möller, I. 2012. Bio-physical linkages in coastal wetlands – implications for coastal protection. Crossing Borders in Coastal Research: Jubilee Conference Proceedings. https://doi.org/10.3990/2.170Google Scholar
Möller, I., Kudella, M., Rupprecht, F., Spencer, T., Paul, M., van Wesenbeeck, B. K., Wolters, G., et al. 2014. Wave attenuation over coastal salt marshes under storm surge conditions. Nature Geoscience, 7: 727731.Google Scholar
Morgan, P. A., Burdick, D. M., and Short, F. T. 2009. The functions and values of fringing salt marshes in northern New England, USA. Estuaries and Coasts, 32: 483495.Google Scholar
Morton, R. A., and Barras, J. A. 2011. Hurricane impacts on coastal wetlands: A half-century record of storm-generated features from Southern Louisiana. Journal of Coastal Research, 27: 2743.Google Scholar
Muller, J., Collins, J. M., Gibson, S. and Paxton, L. 2017. Recent advances in the emerging field of paleotempestology. In: Hurricanes and Climate Change. Springer, Cham, pp. 133.Google Scholar
Nikitina, D. L., Kemp, A. C., Horton, B. P., Vane, C. H., van de Plassche, O., and Engelhart, S. E. 2014. Storm erosion during the past 2000 years along the north shore of Delaware Bay, USA. Geomorphology, 208: 160172.Google Scholar
Oliva, F., Viau, A. E., Peros, M. C., and Bouchard, M. 2018. Paleotempestology database for the western North Atlantic basin. Holocene, 28: 16641671.Google Scholar
Orton, P., Talke, S., Jay, D., Yin, L., Blumberg, A., Georgas, N., Zhao, H., Roberts, H. J., and MacManus, K. 2015. Channel shallowing as mitigation of coastal flooding. Journal of Marine Science and Engineering, 3: 654673.Google Scholar
Pielke, R. Jr, Gratz, J., and Landsea, C. 2008. Normalized hurricane damage in the United States: 1900–2005. Natural Hazards Review, 29–42. https://ascelibrary.org/doi/10.1061/%28ASCE%291527-6988%282008%299%3A1%2829%29Google Scholar
van de Plassche, O., Erkens, G., van Vliet, F., Brandsma, J., van der Borg, K., and de Jong, A. F. M. 2006. Salt-marsh erosion associated with hurricane landfall in southern New England in the fifteenth and seventeenth centuries. Geology, 34: 829832.Google Scholar
van de Plassche, O., van der Borg, K., and de Jong, A. F. M. 1999. Sea level – climate correlation during the past 1400 yr. Geology, 26: 319322.Google Scholar
van de Plassche, O., Wright, A. J., van der Borg, K., and de Jong, A. F. M. 2004. On the erosive trail of a 14th and 15th century hurricane in Connecticut (USA) salt marshes. Radiocarbon, 46: 11111150.Google Scholar
Postma, H. 1961. Transport and accumulation of suspended matter in the Dutch Wadden Sea. Netherlands Journal of Sea Reasearch, 1: 148190.Google Scholar
Priestas, A., Mariotti, G., Leonardi, N., and Fagherazzi, S. 2015. Coupled wave energy and erosion dynamics along a salt marsh boundary, Hog Island Bay, Virginia, USA. Journal of Marine Science and Engineering, 3: 10411065.Google Scholar
Redfield, A. 1972. Development of a New England Salt Marsh. Ecological Monographs, 42: 201237.Google Scholar
Redfield, A. C. 1965. Ontogeny of a salt marsh estuary. Science, 147: 5055.Google Scholar
Resio, D. T., and Westerink, J. J. 2008. Modeling the physics of storm surges. Physics Today, 61: 3338.Google Scholar
Schuerch, M., Vafeidis, A., Slawig, T., and Temmerman, S. 2013. Modeling the influence of changing storm patterns on the ability of a salt marsh to keep pace with sea level rise. Journal of Geophysical Research: Earth Surface, 118: 8496.Google Scholar
Schwimmer, R. A., and Pizzuto, J. E. 2000. A model for the evolution of marsh shorelines. Journal of Sedimentary Research, 70: 10261035.Google Scholar
Shennan, I., and Horton, B. 2002. Holocene land- and sea-level changes in Great Britain. Journal of Quaternary Science, 175–6: 511526.Google Scholar
Shepard, C. C., Crain, C. M., and Beck, M. W. 2011. The protective role of coastal marshes: A systematic review and meta-analysis. PLOS ONE, 6: e27374. https://doi.org/10.1371/journal.pone.0027374Google Scholar
Snedden, G. A., Cretini, K., and Patton, B. 2014. Inundation and salinity impacts to above- and belowground productivity in Spartina patens and Spartina alterniflora in the Mississippi River deltaic plain: Implications for using river diversions as restoration tools. Ecological Engineering, 81: 133139.Google Scholar
Stark, J., Van Oyen, T., Meire, P., and Temmerman, S. 2015. Observations of tidal and storm surge attenuation in a large tidal marsh. Limnology and Oceanography, 60: 13711381.Google Scholar
Stumpf, R. P. 1983. The process of sedimentation on the surface of a salt marsh. Estuarine, Coastal and Shelf Science, 17: 495508.Google Scholar
Tate, A. S., and Battaglia, L. L. 2013. Community disassembly and reassembly following experimental storm surge and wrack application. Journal of Vegetation Science, 24: 4657.Google Scholar
Temmerman, S., Bouma, T. J., Govers, G., Wang, Z. B., De Vries, M. B., and Herman, P. M. J. 2005. Impact of vegetation on flow routing and sedimentation patterns: Three-dimensional modeling for a tidal marsh. Journal of Geophysical Research: Earth Surface, 110: 118.Google Scholar
Temmerman, S., Meire, P., Bouma, T. J., Herman, P. M. J., Ysebaert, T., and De Vriend, H. J. 2013. Ecosystem-based coastal defence in the face of global change. Nature, 504: 7983.Google Scholar
Turner, R. E., Baustian, J. J., Swenson, E. M., and Spicer, J. S. 2006. Wetland sedimentation from hurricanes Katrina and Rita. Science, 314: 449452.Google Scholar
Tweel, A. W., and Turner, R. E. 2012. Landscape-scale analysis of wetland sediment deposition from four tropical cyclone events. PLOS ONE, 7(11). https://doi.org/10.1371/journal.pone.0050528Google Scholar
Walsh, K. J. E., McBride, J. L., Klotzbach, P. J., Balachndran, S., Camargo, S. J., Holland, G. J., … Sugi, M. 2016. Tropical cyclones and climate change. Wiley Interdisciplinary Reviews: Climate Change, 7: 6589.Google Scholar
Walters, D. C., and Kirwan, M. L. 2016. Optimal hurricane overwash thickness for maximizing marsh resilience to sea level rise. Ecology and Evolution, 6: 29482956.Google Scholar
Walters, D., Moore, L. J., Vinent, O. D., Fagherazzi, S., and Mariotti, G. 2014. Interactions between barrier islands and marshes affect island system response to sea level rise: Insights from a coupled model. Journal of Geophysical Research: Earth Surface, 119: 20132031.Google Scholar
Wamsley, T. V., Cialone, M. A., Smith, J. M., Ebersole, B. A., and Grzegorzewski, A. S. 2009. Influence of landscape restoration and degradation on storm surge and waves in southern Louisiana. Natural Hazards, 51: 207224.Google Scholar
Watson, E. B., Wigand, C., Davey, E. W., Andrews, H. M., Bishop, J., and Raposa, K. B. 2017. Wetland Loss Patterns and Inundation-Productivity Relationships Prognosticate Widespread Salt Marsh Loss for Southern New England. Estuaries and Coasts, 40: 662681.Google Scholar
Williams, H. F. L. 2012. Magnitude of Hurricane Ike storm surge sedimentation: Implications for coastal marsh aggradation. Earth Surface Processes and Landforms, 37: 901906.Google Scholar
Wilson, K. R., Kelley, J. T., Croitoru, A., Dionne, M., Belknap, D. F., and Steneck, R. 2009. Stratigraphic and ecophysical characterizations of salt pools: Dynamic landforms of the webhannet salt marsh, wells, ME, USA. Estuaries and Coasts, 32: 855870.Google Scholar

References

Alizad, K., Hagen, S. C., Morris, J. T., Bacopoulos, P., Bilskie, M. V., Weishampel, J. F., and Medeiros, S. C. 2016. A coupled, two-dimensional hydrodynamic-marsh model with biological feedback. Ecological Modelling, 327: 2943.Google Scholar
Allen, J. R. L. 1990. Salt-marsh growth and stratification: A numerical model with special reference to the Severn Estuary, southwest Britain. Marine Geology 95: 7796.Google Scholar
Baptist, M. J., Babovic, V., Uthurburu, J. R., Keijzer, M., Uittenbogaard, R. E., Mynett, A., and Verwey, A. 2007. On inducing equations for vegetation resistance. Journal of Hydraulic Research, 45: 435450.CrossRefGoogle Scholar
Bertness, M. D. 1991. Zonation of Spartina patens and Spartina alterniflora in New England salt marshEcology72: 138148.Google Scholar
Cahoon, D. R., French, J. R., Spencer, T., Reed, D., and Möller, I. 2000. Vertical accretion versus elevational adjustment in UK saltmarshes: An evaluation of alternative methodologies. Geological Society of London Special Publication, 175: 223238.Google Scholar
Cahoon, D. R., Hensel, P. F., Spencer, T., Reed, D. J., McKee, K. L., and Saintilan, N. 2006. Coastal wetland vulnerability to relative sea-level rise: Wetland elevation trends and process controls. In Wetlands and Natural Resource Management. In: Verhoeven, J. T. A., Beltman, ., Bobbink, R., and Whigham, D. F., eds, Berlin, Heidelberg: Springer Berlin Heidelberg. pp. 271292.Google Scholar
Castagno, K. A., Jiménez-Robles, A. M., Donnelly, J. P., Wiberg, P. L., Fenster, M. S. and Fagherazzi, S., 2018. Intense storms increase the stability of tidal baysGeophysical Research Letters, 45: 54915500.Google Scholar
Chen, C., Liu, H., and Beardsley, R. 2003. An unstructured grid, finitevolume, three-dimensional, primitive equations ocean model: Application to coastal ocean and estuaries, Journal of Atmospheric Oceanic Technology, 20: 159186.Google Scholar
Chen, C., Qi, J., Li, C., Beardsley, R. C., Lin, H., Walker, R., and Gates, K., 2008. Complexity of the flooding/drying process in an estuarine tidal‐creek salt‐marsh system: An application of FVCOM. Journal of Geophysical Research: Oceans, 113(C7).Google Scholar
Christiansen, T., Wiberg, P. L., and Milligan, T. G. 2000. Flow and sediment transport on a tidal salt marsh surface. Estuarine and Coastal Shelf Science, 50: 315331.Google Scholar
Chu-Agor, M. L., Muñoz-Carpena, R., Kiker, G., Emanuelsson, A. and Linkov, I., 2011. Exploring vulnerability of coastal habitats to sea level rise through global sensitivity and uncertainty analyses. Environmental Modelling & Software, 26: 593604.Google Scholar
Craft, C., Clough, J., Ehman, J., Joye, S., Park, R., Pennings, S., Guo, H. and Machmuller, M., 2009. Forecasting the effects of accelerated sea‐level rise on tidal marsh ecosystem services. Frontiers in Ecology and the Environment, 7: 7378.CrossRefGoogle Scholar
D'Alpaos, A., Lanzoni, S., Marani, M., Fagherazzi, S. and Rinaldo, A., 2005. Tidal network ontogeny: Channel initiation and early developmentJournal of Geophysical Research: Earth Surface110(F2).CrossRefGoogle Scholar
D’Alpaos, A., Lanzoni, S., Marani, M., and Rinaldo, A. 2007. Landscape evolution in tidal embayments: Modeling the interplay of erosion, sedimentation, and vegetation dynamics. Journal of Geophysical Research: Earth Surface 112 (F1). Wiley Online Library.Google Scholar
D’Alpaos, A., and Marani, M. 2016. Reading the signatures of biologic–geomorphic feedbacks in salt-marsh landscapes. Advances in Water Resources, 93: 265275.Google Scholar
Einstein, H. A., and Krone, R. B. 1962. Experiments to determine modes of cohesive sediment transport in salt water, Journal of Geophysical Research, 67: 14511461.Google Scholar
Fagherazzi, S., Kirwan, M. L., Mudd, S. M., Guntenspergen, G. R., Temmerman, S., D'Alpaos, A., Koppel, J., et al. 2012. Numerical models of salt marsh evolution: Ecological, geomorphic, and climatic factors. Reviews of Geophysics, 50(1): 128.Google Scholar
Fagherazzi, S., Marani, M., and Blum, L.K. (eds) 2004. The Ecogeomorphology of Tidal Marshes, Vol. 59, Coastal and Estuarine Studies. American Geophysical Union, Washington, D.C.Google Scholar
French, J. R. 1993. Numerical simulation of vertical marsh growth and adjustment to accelerated sea-level rise, North Norfolk, U.K. Earth Surface Processes and Landforms 18: : 6381.Google Scholar
Goldenfeld, N., and Kadanoff, L. P. 1999. Simple lessons from complexity. Science, 284: 8789.Google Scholar
Kirwan, M. L., Guntenspergen, G. R., D’Alpaos, A., Morris, J. T., Mudd, S. M., and Temmerman, S. 2010. Limits on the adaptability of coastal marshes to rising sea level. Geophysical Research Letters, 37: 15.Google Scholar
Kirwan, M. L., and Murray, A. B. 2007. A coupled geomorphic and ecological model of tidal marsh evolution. Proceedings of the National Academy of Sciences of the United States of America, 104: 6118–22.Google Scholar
Kirwan, M. L., and Temmerman, S. 2009. Coastal marsh response to historical and future sea-level acceleration. Quaternary Science Reviews, 28: 1801–8.Google Scholar
Kirwan, M. L., Walters, D. C., Reay, W. G., and Carr, J. A. 2016. Sea level driven marsh expansion in a coupled model of marsh erosion and migration. Geophysical Research Letters, 43: 43664373.Google Scholar
Leonardi, N., and Fagherazzi, S. 2014. How waves shape salt marshes. Geology, 42: 887890.Google Scholar
Leonardi, N., and Fagherazzi, S. 2015. Effect of local variability in erosional resistance on large‐scale morphodynamic response of salt marshes to wind waves and extreme events. Geophysical Research Letters, 42: 58725879.Google Scholar
Leonardi, N., Ganju, N. K., and Fagherazzi, S. 2016. A linear relationship between wave power and erosion determines salt-marsh resilience to violent storms and hurricanes. Proceedings of the National Academy of Sciences of the United States of America, 113: 6468.Google Scholar
Lesser, G., Roelvink, J., Van Kester, J., and Stelling, G. 2004. Development and validation of a three-dimensional morphological model. Coastal Engineering, 51: 883915.Google Scholar
Lorenzo-Trueba, J., and Ashton, A. D. 2014. Rollover, drowning, and discontinuous retreat: Distinct modes of barrier response to sea-level rise arising from a simple morphodynamic model. Journal of Geophysical Research: Earth Surface, 119: 779801.Google Scholar
Lorenzo-Trueba, J., and Mariotti, G. 2017. Chasing boundaries and cascade effects in a coupled barrier-marsh-lagoon system. Geomorphology, 290: 153163.Google Scholar
Luettich, R. A., Westerink, J. J., and Scheffner, N. W. 1992. ADCIRC: An advanced three-dimensional circulation model for shelves, coasts, and estuaries. I: Theory and methodology of ADCIRC-2DD1 and ADCIRC-3DL. In: Technical Rep. No. DRP-92-6. U.S. Army Engineer Waterways Experiment Station, Vicksburg, MS.Google Scholar
Marani, M., D’Alpaos, A., Lanzoni, S., Carniello, L., and Rinaldo, A. 2007. Biologically-controlled multiple equilibria of tidal landforms and the fate of the Venice Lagoon. Geophysical Research Letters, 34: L11402.Google Scholar
Marani, M., Lio, C. D., and D’Alpaos, A. 2013. Vegetation engineers marsh morphology through multiple competing stable states. Proceedings of the National Academy of Sciences of the United States of America, 110: 32593263.Google Scholar
Mariotti, G., and Carr, J. 2014. Dual role of salt marsh retreat: Long-term loss and short-term resilience. Water Resources Research, 50: 29632974.Google Scholar
Mariotti, G., and Fagherazzi, S. 2010. A numerical model for the coupled long-term evolution of salt marshes and tidal flats. Journal of Geophysical Research: Earth Surface,. 115: F01004.Google Scholar
Mariotti, G. and Fagherazzi, S. 2013. Critical width of tidal flats triggers marsh collapse in the absence of sea-level rise. Proceedings of the National Academy of Sciences of the United States of America, 110: 53535356.Google Scholar
Mehta, A. J. 1984. Characterization of cohesive sediment properties and transport processes in estuaries. In: Mehta, A. J., ed., Estuarine Cohesive Sediment Dynamics, Lecture Notes on Coastal and Estuarine Studies, vol. 14, Springer, New York, pp. 290315.Google Scholar
Mendez, F. J. and Losada, I. J. 2004. An empirical model to estimate the propagation of random breaking and nonbreaking waves over vegetation fieldsCoastal Engineering51: 103118.Google Scholar
Morris, J. T., Sundareshwar, P. V., Nietch, C. T., Kjerfve, B., and Cahoon, D. R. 2002. Responses of coastal wetlands to rising sea level, Ecology, 83: 28692877.Google Scholar
Mudd, S. M., Fagherazzi, S., Morris, J. T., and Furbish, D. J. 2004. Flow, sedimentation, and biomass production on a vegetated salt marsh in South Carolina: Toward a predictive model of marsh morphologic and ecologic evolution. In: Fagherazzi, S., Marani, M., and Blum, L. K., eds., The Ecogeomorphology of Salt Marshes, Coastal Estuarine Studies, vol. 59, American Geophysical Union, Washington, DC, pp. 165188.Google Scholar
Mudd, S. M., Howell, S. M. and Morris, J. T. 2009. Impact of dynamic feedbacks between sedimentation, sea-level rise, and biomass production on near-surface marsh stratigraphy and carbon accumulation. Estuarine, Coastal and Shelf Science, 82: 377389.Google Scholar
Murray, A. B. 2003. Contrasting the goals, strategies, and predictions associated with simplified numerical models and detailed simulations. In: Wilcock, P. R. and Iverson, R. M., eds, Prediction in Geomorphology, American Geophysical Union, Washington DC, pp. 151165.Google Scholar
Murray, A. B., Gasparini, N. M., Goldstein, E. B., and van der Wegen, M. 2016. Uncertainty quantification in modeling earth surface processes: More applicable for some types of models than for others. Computers & Geosciences, 90: 616.Google Scholar
Nardin, W. and Edmonds, D. A. 2014. Optimum vegetation height and density for inorganic sedimentation in deltaic marshes. Nature Geoscience, 7: 722726.Google Scholar
Nardin, W., Edmonds, D. A., and Fagherazzi, S. 2016. Influence of vegetation on spatial patterns of sediment deposition in deltaic islands during floodAdvances in Water Resources93: 236248.Google Scholar
Nardin, W., Larsen, L., Fagherazzi, S., and Wiberg, P., 2018. Tradeoffs among hydrodynamics, sediment fluxes and vegetation community in the Virginia Coast Reserve, USAEstuarine, Coastal and Shelf Science, 210: 98108.CrossRefGoogle Scholar
Nepf, H. M., 1999. Drag, turbulence, and diffusion in flow through emergent vegetationWater Resources Research35: 479489.Google Scholar
Neumeier, U., 2005. Quantification of vertical density variations of salt-marsh vegetationEstuarine, Coastal and Shelf Science63: 489496.Google Scholar
Priestas, A. M., Mariotti, G., Leonardi, N., and Fagherazzi, S., 2015. Coupled wave energy and erosion dynamics along a salt marsh boundary, Hog Island Bay, Virginia, USA. Journal of Marine Science and Engineering, 3: 10411065.Google Scholar
Rahmstorf, S. 2007. A semi-empirical approach to projecting future sea-level rise. Science, 315: 368370.CrossRefGoogle ScholarPubMed
Ratliff, K. M., Braswell, A. E., and Marani, M. 2015. Spatial response of coastal marshes to increased atmospheric CO2. Proceedings of the National Academy of Sciences of the United States of America, 112, 1558015584.CrossRefGoogle ScholarPubMed
Redfield, A. C. 1965. Ontogeny of a salt marsh estuary. Science 147: 5055.Google Scholar
Rinaldo, A., Fagherazzi, S., Lanzoni, S., Marani, M., and Dietrich, W. E. 1999. Tidal networks: 2. Watershed delineation and comparative network morphologyWater Resources Research35: 39053917.Google Scholar
Rodi, W., 1980. Turbulence Models and Their Application in Hydraulics. International Association for Hydro-Environment Engineering and Research, Delft.Google Scholar
Schwimmer, R. A., 2001. Rates and processes of marsh shoreline erosion in Rehoboth Bay, Delaware, USAJournal of Coastal Research, 17: 672683.Google Scholar
Stolper, D., List, J. H., and Thieler, E. R. 2005. Simulating the evolution of coastal morphology and stratigraphy with a new morphological-behaviour model (GEOMBEST). Marine Geology, 218: 1736.Google Scholar
Tambroni, N., and Seminara, G. 2012. A one-dimensional eco-geomorphic model of marsh response to sea level rise: Wind effects, dynamics of the marsh border and equilibrium. Journal of Geophysical Research: Earth Surface, 117: F03026.Google Scholar
Temmerman, S., Bouma, T. J., Govers, G., Wang, Z. B., De Vries, M. B., and Herman, P. M. J. 2005. Impact of vegetation on flow routing and sedimentation patterns: Three‐dimensional modeling for a tidal marsh. Journal of Geophysical Research: Earth Surface, 110: F04019.Google Scholar
Temmerman, S., Bouma, T. J., Van de Koppel, J., Van der Wal, D., De Vries, M. B. and Herman, P. M. J. 2007. Vegetation causes channel erosion in a tidal landscape. Geology, 35: 631634.Google Scholar
Temmerman, S., Govers, G., Meire, P., and Wartel, S. 2003. Modelling long-term tidal marsh growth under changing tidal conditions and suspended sediment concentrations, Scheldt estuary, Belgium. Marine Geology 193: 151169.Google Scholar
Temmerman, S., Govers, G., Meire, P., and Wartel, S. 2004. Simulating the long-term development of levee–basin topography on tidal marshes. Geomorphology, 63: 3955.Google Scholar
Walters, D., Moore, L. J., Duran Vinent, O., Fagherazzi, S., and Mariotti, G. 2014. Interactions between barrier islands and backbarrier marshes affect island system response to sea level rise: Insights from a coupled model. Journal of Geophysical Research: Earth Surface, 119: F003091.Google Scholar
Weston, N. B. 2014. Declining sediments and rising seas: An unfortunate convergence for tidal wetlandsEstuaries and Coasts37: 123.Google Scholar
Zhao, L., Chen, C., Vallino, J., Hopkinson, C., Beardsley, R. C., Lin, H., and Lerczak, J., 2010. Wetland‐estuarine‐shelf interactions in the Plum Island Sound and Merrimack River in the Massachusetts coast. Journal of Geophysical Research: Oceans, 115: C10039.Google Scholar
Zong, L. and Nepf, H., 2010. Flow and deposition in and around a finite patch of vegetationGeomorphology116: 363372.Google Scholar

References

JNCC. An Overview of Coastal Saltmarshes, Their Dynamic and Sensitivity Characteristics for Conservation and Management. UK: JNCC, 2003.Google Scholar
Tempest, JA, Möller, I, Spencer, T. A review of plant-flow interactions on salt marshes: The importance of vegetation structure and plant mechanical characteristics: Salt marsh plant-flow interactions. Wiley Interdiscip Rev Water 2015;2: 669681.Google Scholar
Townend, I, Fletcher, C, Knappen, M et al. A review of salt marsh dynamics: A review of salt marsh dynamics. Water Environ J 2011;25: 477488.Google Scholar
van Wesenbeeck, BK, van de Koppel, J, Herman, PMJ et al. Does scale-dependent feedback explain spatial complexity in salt-marsh ecosystems? Oikos 2008;117: 152159.Google Scholar
Wang, H, van der Wal, D, Li, X et al. Zooming in and out: Scale dependence of extrinsic and intrinsic factors affecting salt marsh erosion: Factors on salt marsh edge erosion. J Geophys Res Earth Surf 2017;122: 14551470.Google Scholar
Reise, K. Sediment mediated species interactions in coastal waters. J Sea Res 2002;48: 127141.Google Scholar
Shi, Z, Pethick, JS, Pye, K. Flow structure in and above the various heights of a saltmarsh canopy: a laboratory flume study. J Coast Res 1995: 1204–1209.Google Scholar
Temmerman, S, Bouma, TJ, Govers, G et al. Impact of vegetation on flow routing and sedimentation patterns: Three-dimensional modeling for a tidal marsh: Vegetation impact on flow routing. J Geophys Res Earth Surf 2005;110, DOI: 10.1029/2005JF000301.Google Scholar
Chen, Y, Li, Y, Cai, T et al. A comparison of biohydrodynamic interaction within mangrove and saltmarsh boundaries: Bio-hydrodynamics within mangrove and saltmarsh boundaries. Earth Surf Process Landforms 2016;41: 19671979.Google Scholar
Möller, I, Spencer, T, French, JR et al. Wave transformation over salt marshes: A field and numerical modelling study from North Norfolk, England. Estuarine, Coast Shelf Sci 1999;49: 411426.Google Scholar
Neumeier, U. Velocity and turbulence variations at the edge of saltmarshes. Continental Shelf Res 2007;27: 10461059.Google Scholar
Shi, B, Wang, YP, Yang, Y et al. Determination of critical shear stresses for erosion and deposition based on in situ measurements of currents and waves over an intertidal mudflat. J Coast Res 2015;316: 13441356.Google Scholar
Lo, V, Bouma, T, van Colen, C, Airoldi, L. Interactive effects of vegetation and grain size on erosion rates in salt marshes of the Northern Adriatic Sea. ESCA Local Meeting report. 2016.Google Scholar
Reed, DJ. Sediment dynamics and deposition in a retreating coastal salt marsh. Estuarine, Coast Shelf Sci 1988;26: 6779.Google Scholar
Shi, Y, Jiang, B, Nepf, HM. Influence of particle size and density, and channel velocity on the deposition patterns around a circular patch of model emergent vegetation patch on deposition patterns. Water Resour Res 2016;52: 1044–1055.Google Scholar
Alben, S, Shelley, M, Zhang, J. Drag reduction through self-similar bending of a flexible body. Nature 2002;420: 479481.Google Scholar
Folkard, AM. Hydrodynamics of model Posidonia oceanica patches in shallow water. Limnol Oceanogr 2005;50: 15921600.Google Scholar
Cowell, PJ, Thom, BG. Morphodynamics of coastal evolution. In: Carter, W, Woodroffe, CD. (eds.). Coastal Evolution : Late Quaternary Shoreline Morphodynamics Cambridge University Press, Cambridge, 1994, 3386.Google Scholar
Yang, SL, Shi, BW, Bouma, TJ et al. Wave attenuation at a salt marsh margin: A Case study of an exposed coast on the Yangtze Estuary. Estuaries Coasts 2012;35: 169182.Google Scholar
Ysebaert, T, Yang, S-L, Zhang, L et al. Wave attenuation by two contrasting ecosystem engineering salt marsh macrophytes in the intertidal pioneer zone. Wetlands 2011;31: 10431054.Google Scholar
Shi, Z, Pethick, JS, Burd, F et al. Velocity profiles in a salt marsh canopy. Geo-Marine Lett 1996;16: 319323.Google Scholar
Neumeier, U, Amos, CL. The influence of vegetation on turbulence and flow velocities in European salt-marshes. Sedimentology 2006;53: 259277.Google Scholar
Leonard, LA, Croft, AL. The effect of standing biomass on flow velocity and turbulence in Spartina alterniflora canopies. Estuarine, Coast Shelf Sci 2006;69: 325336.Google Scholar
Leonard, LA and Reed, DJ. Hydrodynamics and sediment transport through tidal marsh canopies. Journal of Coastal Research 2017; SI36: 459469.Google Scholar
Neumeier, U, Ciavola, P. Flow resistance and associated sedimentary processes in a Spartina maritima salt-marsh. J Coast Res 2004;202: 435447.Google Scholar
Neumeier, U. Quantification of vertical density variations of salt-marsh vegetation. Estuarine, Coast Shelf Sci 2005;63: 489496.Google Scholar
Pethick, J, Leggett, D, Husain, L. Boundary layers under salt marsh vegetation developed in tidal currents. In: Thornes, J. (ed.). Vegetation and Erosion. John Wiley and Sons, Chichester, 1990, 113124.Google Scholar
Tollner, E., Barfield, B., Hayes, J. Sedimentology of erect vegetal filters. J. Hydraulics Division 1982;108: 15181531.Google Scholar
Möller, I, Kudella, M, Rupprecht, F et al. Wave attenuation over coastal salt marshes under storm surge conditions. Nat Geosci 2014;7: 727731.Google Scholar
Bouma, TJ, De Vries, MB, Herman, PMJ. Comparing ecosystem engineering efficiency of two plant species with contrasting growth strategies. Ecology 2010;91: 26962704.Google Scholar
Tempest, J. Hydrodynamic effects of salt marsh canopies and their prediction using remote sensing techniques. 2017. PhD Thesis. Cambridge University.Google Scholar
Horn, R, Richards, K. Flow–vegetation interactions in restored floodplain environments. Hydroecology and Ecohydrology. John Wiley & Sons, Ltd, Chichester, 2008, 269294.Google Scholar
Nepf, HM, Vivoni, ER. Flow structure in depth‐limited, vegetated flow. J Geophys Res Ocean 2000;105: 2854728557.Google Scholar
Lopez, F, Garcia, M. Open-channel flow through simulated vegetation: Suspended sediment transport modeling. Water Resour Res 1998;34: 23412352.Google Scholar
Anderson, ME, Smith, JM. Wave attenuation by flexible, idealized salt marsh vegetation. Coast Eng 2014;83: 8292.Google Scholar
Stapleton, KR, Huntley, DA. Seabed stress determinations using the inertial dissipation method and the turbulent kinetic energy method. Earth Surf Process Landforms, 20: 807–815.Google Scholar
Wilson, CAME, Stoesser, T, Bates, PD. Modelling of open channel flow through vegetation. Computational Fluid Dynamics: Applications in Environmental Hydraulics. John Wiley & Sons, Chichester, 2005, 395428.Google Scholar
Leonard, LA, Luther, ME. Flow hydrodynamics in tidal marsh canopies. Limnol Oceanogr 1995;40: 14741484.Google Scholar
Christiansen, T, Wiberg, PL, Milligan, TG. Flow and sediment transport on a tidal salt marsh surface. Estuarine, Coast Shelf Sci 2000;50: 315331.Google Scholar
Lightbody, AF, Nepf, HM. Prediction of velocity profiles and longitudinal dispersion in salt marsh vegetation. Limnol Oceanogr 2006;51: 218228.Google Scholar
Paulus, S, Schumann, H, Kuhlmann, H, Léon, J. High-precision laser scanning system for capturing 3D plant architecture and analysing growth of cereal plants. Biosyst Eng 2014;121: 111.Google Scholar
Luhar, M, Nepf, HM. Flow-induced reconfiguration of buoyant and flexible aquatic vegetation. Limnol Oceanogr 2011;56: 20032017.Google Scholar
Bradley, K, Houser, C. Relative velocity of seagrass blades: Implications for wave attenuation in low-energy environments. J Geophys Res 2009;114, DOI: 10.1029/2007JF000951.Google Scholar
Albayrak, I, Nikora, V, Miler, O, O’Hare, M. Flow-plant interactions at a leaf scale: Effects of leaf shape, serration, roughness and flexural rigidity. Aquat Sci 2012;74: 267286.CrossRefGoogle Scholar
Feagin, RA, Irish, JL, Möller, I et al. Short communication: Engineering properties of wetland plants with application to wave attenuation. Coast Eng 2011;58: 251255.Google Scholar
Rupprecht, F, Möller, I, Evans, B et al. Biophysical properties of salt marsh canopies — Quantifying plant stem flexibility and above ground biomass. Coast Eng 2015;100: 4857.Google Scholar
Lagerspetz, KY. Thermal acclimation without heat shock, and motor responses to a sudden temperature change in Asellus aquaticus. J Therm Biol 2003;28: 421427.Google Scholar
Díaz, S, Symstad, AJ, Stuart Chapin, F et al. Functional diversity revealed by removal experiments. Trends Ecol & Evol 2003;18: 140146.Google Scholar
Paul, M, Bouma, T, Amos, C. Wave attenuation by submerged vegetation: Combining the effect of organism traits and tidal current. Mar Ecol Prog Ser 2012;444: 3141.Google Scholar
Bouma, TJ, van Duren, LA, Temmerman, S et al. Spatial flow and sedimentation patterns within patches of epibenthic structures: Combining field, flume and modelling experiments. Cont Shelf Res 2007;27: 10201045.Google Scholar
Panigrahi, K, Khatua, KK. Prediction of velocity distribution in straight channel with rigid vegetation. Aquat Procedia 2015;4: 819825.Google Scholar
Kobayashi, N, Raichle, AW, Asano, T. Wave attenuation by vegetation. J Waterw Port, Coastal, Ocean Eng 1993;119: 3048.Google Scholar
John, BM, Shirlal, KG, Rao, S. Effect of artificial sea grass on wave attenuation- an experimental investigation. Aquat Procedia 2015;4: 221226.Google Scholar
John, BM, Shirlal, KG, Rao, S et al. Effect of artificial seagrass on wave attenuation and wave run-up. Int J Ocean Clim Syst 2016;7, DOI: 10.1177/1759313115623163.Google Scholar
Stratigaki, V, Manca, E, Prinos, P et al. Large-scale experiments on wave propagation over Posidonia oceanica. J Hydraul Res 2011;49: 3143.Google Scholar
Hu, Z, Suzuki, T, Zitman, T et al. Laboratory study on wave dissipation by vegetation in combined current–wave flow. Coast Eng 2014;88: 131142.Google Scholar
Okamoto, T-A, Nezu, I. Turbulence structure and “Monami” phenomena in flexible vegetated open-channel flows. J Hydraul Res 2009;47: 798810.Google Scholar
Murphy, E, Ghisalberti, M, Nepf, H. Model and laboratory study of dispersion in flows with submerged vegetation. Water Resour Res 2007;43. https://doi.org/10.1029/2006WR005229Google Scholar
Wu, F-C, Shen, HW, Chou, Y-J. Variation of roughness coefficients for unsubmerged and submerged vegetation. J Hydraul Eng 1999;125: 934–42.Google Scholar
Stone, BM, Shen, HT. Hydraulic resistance of flow in channels with cylindrical roughness. J Hydraul Eng 2002;128: 500506.Google Scholar
Augustin, LN, Irish, JL, Lynett, P. Laboratory and numerical studies of wave damping by emergent and near-emergent wetland vegetation. Coast Eng 2009;56: 332340.Google Scholar
Fonseca, MS, Koehl, MAR. Flow in seagrass canopies: The influence of patch width. Estuarine, Coast Shelf Sci 2006;67: 19.Google Scholar
Folkard, AM. Flow regimes in gaps within stands of flexible vegetation:Laboratory flume simulations. Environ Fluid Mech 2011;11: 289306.Google Scholar
Wilson, CAME, Stoesser, T, Bates, PD, Stoesser, T. Open channel flow through different forms of submerged flexible vegetation. J Hydraul Eng 2003;129: 847853.Google Scholar
Ghisalberti, M, Nepf, H. The structure of the shear layer in flows over rigid and flexible canopies. Environ Fluid Mech 2006;6: 277301.Google Scholar
Elliott, AH. Settling of fine sediment in a channel with emergent vegetation. J Hydraul Eng 2000;126: 570577.Google Scholar
Shi, Y, Jiang, B, Nepf, HM. Influence of particle size and density, and channel velocity on the deposition patterns around a circular patch of model emergent vegetation. Water Resour Res 2016; 52: 10441055.Google Scholar
Bouma, TJ, De Vries, MB, Low, E et al. Trade‐offs related to ecosystem engineering: A case study on stiffness of emerging macrophytes. Ecology 2005;86: 21872199.Google Scholar
Eckman, JE, Nowell, ARM. Boundary skin friction and sediment transport about an animal-tube mimic. Sedimentology 1984;31: 851862.Google Scholar
Eckman, JE. Flow disruption by an animal-tube mimic affects sediment bacterial colonization. J Mar Res 1985;43: 419435.Google Scholar
Wilson, CAME, Yagci, O, Rauch, H ‐P, Stoesser, T. Application of the drag force approach to model the flow‐interaction of natural vegetation. Int J River Basin Manag 2006;4: 137–46.Google Scholar
Paul, M, Henry, P-YT, Thomas, RE. Geometrical and mechanical properties of four species of northern European brown macroalgae. Coast Eng 2014;84: 7380.Google Scholar
Anderson, ME, Smith, JM. Wave attenuation by flexible, idealized salt marsh vegetation. Coast Eng 2014;83: 8292.Google Scholar
John, BM, Shirlal, KG, Rao, S. Effect of artificial vegetation on wave attenuation – an experimental investigation. Procedia Eng 2015;116: 600606.Google Scholar
Loder, NM, Irish, JL, Cialone, MA, Wamsley, TV. Sensitivity of hurricane surge to morphological parameters of coastal wetlands. Estuarine, Coast Shelf Sci 2009;84: 625636.Google Scholar
Spencer, T, Brooks, SM, Evans, BR et al. Southern North Sea storm surge event of 5 December 2013: Water levels, waves and coastal impacts. Earth-Science Rev 2015;146: 120145.Google Scholar
Spencer, T, Möller, I, Rupprecht, F et al. Salt marsh surface survives true-to-scale simulated storm surges. Earth Surf Process Landforms 2016;41: 543552.Google Scholar
Maza, M, Lara, JL, Losada, IJ et al. Large-scale 3-D experiments of wave and current interaction with real vegetation. Part 2: Experimental analysis. Coast Eng 2015;106: 7386.Google Scholar
van Duren, LA, Herman, PMJ, Sandee, AJJ, Heip, CHR. Effects of mussel filtering activity on boundary layer structure. J Sea Res 2006;55: 314.Google Scholar
Carey, DA. Particle resuspension in the benthic boundary layer induced by flow around polychaete tubes. Can J Fish Aquat Sci 1983;40: s301s308.Google Scholar
Gerbersdorf, SU, Wieprecht, S. Biostabilization of cohesive sediments: Revisiting the role of abiotic conditions, physiology and diversity of microbes, polymeric secretion, and biofilm architecture. Geobiology 2015;13: 6897.Google Scholar
Decho, A. Microbial exopolymer secretions in ocean environments: Their role(s) in food webs and marine processes. In: Barnes, M (ed.). Oceanography and Marine Biology: An Annual Review. Aberdeen University Press, 1990.Google Scholar
Andersen, TJ. Seasonal variation in erodibility of two temperate, microtidal mudflats. Estuarine, Coast Shelf Sci 2001;53: 112.Google Scholar
Graf, G, Rosenberg, R. Bioresuspension and biodeposition: A review. J Mar Syst 1997;11: 269278.Google Scholar
Passarelli, C, Olivier, F, Paterson, DM et al. Organisms as cooperative ecosystem engineers in intertidal flats. J Sea Res 2014;92: 92101.Google Scholar
Grant, J, Daborn, G. The effects of bioturbation on sediment transport on an intertidal mudflat. Neth J Sea Res 1994;32: 6372.Google Scholar
Tolhurst, T., Riethmüller, R, Paterson, D. In situ versus laboratory analysis of sediment stability from intertidal mudflats. Cont Shelf Res 2000;20: 13171334.Google Scholar
Droppo, IG, Lau, YL, Mitchell, C. The effect of depositional history on contaminated bed sediment stability. Sci Total Environ 2001;266: 713.Google Scholar
Droppo, IG, Ross, N, Skafel, M, Liss, SN. Biostabilization of cohesive sediment beds in a freshwater wave-dominated environment. Limnol Oceanogr 2007;52: 577589.Google Scholar
Van Colen, C, Underwood, GJC, Serôdio, J, Paterson, DM. Ecology of intertidal microbial biofilms: Mechanisms, patterns and future research needs. J Sea Res 2014;92: 25.Google Scholar
Ubertini, M, Lefebvre, S, Rakotomalala, C, Orvain, F. Impact of sediment grain-size and biofilm age on epipelic microphytobenthos resuspension. J Exp Mar Biol Ecol 2015;467: 5264.Google Scholar
Widdows, J, Brown, S, Brinsley, M. et al. Temporal changes in intertidal sediment erodability: influence of biological and climatic factors. Cont Shelf Res 2000;20: 12751289.Google Scholar
Lundkvist, M, Grue, M, Friend, PL, Flindt, MR. The relative contributions of physical and microbiological factors to cohesive sediment stability. Cont Shelf Res 2007;27: 11431152.Google Scholar
Widdows, J, Friend, PL, Bale, AJ et al. Inter-comparison between five devices for determining erodability of intertidal sediments. Cont Shelf Res 2007;27: 11741189.Google Scholar
Black, KS, Sun, H, Craig, G et al. Incipient erosion of biostabilized sediments examined using particle-field optical holography. Environ Sci & Technol 2001;35: 22752281.Google Scholar
Neumeier, U, Lucas, CH, Collins, M. Erodibility and erosion patterns of mudflat sediments investigated using an annular flume. Aquat Ecol 2006;40: 543554.Google Scholar
da S. Quaresma, V, Amos, CL, Flindt, M. The influences of biological activity and consolidation time on laboratory cohesive beds. J Sediment Res 2004;74: 15271404.Google Scholar
Madsen, KN, Nilsson, P, Sundbäck, K. The influence of benthic microalgae on the stability of a subtidal sediment. J Exp Mar Biol Ecol 1993;170: 159177.Google Scholar
Wolf, G, Picioreanu, C, van Loosdrecht, MCM. Kinetic modeling of phototrophic biofilms: the PHOBIA model. Biotechnol Bioeng 2007;97: 10641079.Google Scholar
Stone, M, Emelko, MB, Droppo, IG, Silins, U. Biostabilization and erodibility of cohesive sediment deposits in wildfire-affected streams. Water Res 2011;45: 521534.Google Scholar
Pacepavicius, G, Lau, YL, Liu, D et al. A rapid biochemical method for estimating biofilm mass. Environ Toxicol Water Qual 1997;12: 97100.Google Scholar
DuBois, M, Gilles, KA, Hamilton, JK et al. Colorimetric method for determination of sugars and related substances. Anal Chem 1956;28, DOI: 10.1021/ac60111a017.Google Scholar
Kochert, AG. Carbohydrate determination by the phenol–sulfuric acid method. In: Hellebust, JA, Craigie, JS (eds.), Handbook of Phycological Methods: Physiological and Biochemical Methods. Cambridge: Cambridge University Press, 1978, 9597.Google Scholar
Bradford, MM. A rapid and sensitive method for the quantitation of microgram quantities of protein utilizing the principle of protein-dye binding. Anal Biochem 1975;72: 248–54.Google Scholar
Lorenzen, CJ. Determination of chlorophyll and pheo-pigments: Spectrophotometric equations. Limnol Oceanogr 1967;12: 343346.Google Scholar
White, DC, Davis, WM, Nickels, JS et al. Determination of the sedimentary microbial biomass by extractible lipid phosphate. Oecologia 1979;40: 5162.Google Scholar
Eaton, J W., Moss, B. The estimation of numbers and pigment content in epipelic algal populations: Estimation of epipelic algal populations. Limnol Oceanogr 1966;11: 584595.Google Scholar
Nowell, A. Flow environments of aquatic benthos. Annu Rev Ecol Syst 1984;15: 303328.Google Scholar
Gerdol, V, Hughes, R. Effect of Corophium volutator on the abundance of benthic diatoms, bacteria and sediment stability in two estuaries in southeastern England. Mar Ecol Prog Ser 1994;114: 109115.Google Scholar
Mazik, K, Elliott, M. The effects of chemical pollution on the bioturbation potential of estuarine intertidal mudflats. Helgol Mar Res 2000;54: 99109.Google Scholar
Le Hir, P, Monbet, Y, Orvain, F. Sediment erodability in sediment transport modelling: Can we account for biota effects? Cont Shelf Res 2007;27: 11161142.Google Scholar
Fernandes, S, Sobral, P, Costa, MH. Nereis diversicolor effect on the stability of cohesive intertidal sediments. Aquat Ecol 2006;40: 567579.Google Scholar
Widdows, J, Brinsley, M, Pope, N. Effect of Nereis diversicolor density on the erodability of estuarine sediment. Mar Ecol Prog Ser 2009;378: 135143.Google Scholar
Grabowski, RC, Droppo, IG, Wharton, G. Erodibility of cohesive sediment: The importance of sediment properties. Earth-Science Rev 2011;105, 101120.Google Scholar
Tolhurst, TJ, Chapman, MG, Underwood, AJ et al. Technical Note: The effects of five different defaunation methods on biogeochemical properties of intertidal sediment. Biogeosciences 2012;9: 36473661.Google Scholar
Hale, R, Jacques, RO, Tolhurst, TJ. Cryogenic defaunation of sediments in the field. J Coast Res 2015;316: 15371540.Google Scholar
Gamenick, I, Jahn, A, Vopel, K et al. Hypoxia and sulphide as structuring factors in a macrozoobenthic community on the Baltic Sea shore: colonisation studies and tolerance experiments. Mar Ecol Prog Ser 1996;144, 7385.Google Scholar
Thrush, SF, Whitlatch, RB, Pridmore, RD et al. Scale-dependent recolonization: The role of sediment stability in a dynamic sandflat habitat. Ecology 1996;77: 24722487.Google Scholar
Kristensen, E, Blackburn, TH. The fate of organic carbon and nitrogen in experimental marine sediment systems: Influence of bioturbation and anoxia. J Mar Res 1987;45: 231257.Google Scholar
Gilbert, F, Rivet, L, Bertrand, J-C. The in vitro influence of the burrowing polychaete Nereis diversicolor on the fate of petroleum hydrocarbons in marine sediments. Chemosphere 1994;29: 112.Google Scholar
Nowell, AR., Jumars, PA, Eckman, JE. Effects of biological activity on the entrainment of marine sediments. Mar Geol 1981;42: 133153.Google Scholar
Orvain, F, Sauriau, P, Sygut, A et al. Interacting effects of Hydrobia ulvae bioturbation and microphytobenthos on the erodibility of mudflat sediments. Mar Ecol Prog Ser 2004;278: 205223.Google Scholar
Needham, HR, Pilditch, CA, Lohrer, AM et al. Density and habitat dependent effects of crab burrows on sediment erodibility. J Sea Res 2013;76: 94104.Google Scholar
Thompson, CEL, Williams, JJ, Metje, N et al. Turbulence based measurements of wave friction factors under irregular waves on a gravel bed. Coast Eng 2012;63: 3947.Google Scholar
Wheatcroft, RA. Temporal variation in bed configuration and one-dimensional bottom roughness at the mid-shelf STRESS site. Cont Shelf Res 1994;14: 11671190.Google Scholar
Lyons, AP, Fox, WLJ, Hasiotis, T et al. Characterization of the two-dimensional roughness of wave-rippled sea floors using digital photogrammetry. IEEE J Ocean Eng 2002;27: 515524.Google Scholar
Briggs, KB. Microtopographical roughness of shallow-water continental shelves. IEEE J Ocean Eng 1989;14: 360367.Google Scholar
Jackson, DR, Briggs, KB. High‐frequency bottom backscattering: Roughness versus sediment volume scattering. J Acoust Soc Am 1992;92, DOI: 10.1121/1.403966.Google Scholar
Widdows, J, Brinsley, M. Impact of biotic and abiotic processes on sediment dynamics and the consequences to the structure and functioning of the intertidal zone. J Sea Res 2002;48: 143156.Google Scholar
Widdows, J, Brinsley, M, Salkeld, P, Lucas, CH. Influence of biota on spatial and temporal variation in sediment erodability and material flux on a tidal flat (Westerschelde, The Netherlands). Mar Ecol Prog Ser 2000;194: 2337.Google Scholar
Escapa, M, Perillo, GME, Iribarne, O. Sediment dynamics modulated by burrowing crab activities in contrasting SW Atlantic intertidal habitats. Estuarine, Coast Shelf Sci 2008;80: 777780.Google Scholar
Widdows, J, Brinsley, MD, Bowley, N, Barrett, C. A benthic annular flume for in situ measurement of suspension feeding/biodeposition rates and erosion potential of intertidal cohesive sediments. Estuarine, Coast Shelf Sci 1998;46: 2738.Google Scholar
Grant, J, Gust, G. Prediction of coastal sediment stability from photopigment content of mats of purple sulphur bacteria. Nature 1987;330: 244246.Google Scholar
Gust, G. Skin friction probes for field applications. J Geophys Res 1988;93: 1412114132.Google Scholar
Blanchard, G, Sauriau, P, Cariou-Le Gall, V et al. Kinetics of tidal resuspension of microbiota: Testing the effects of sediment cohesiveness and bioturbation using flume experiments. Mar Ecol Prog Ser 1997;151: 1725.Google Scholar
Thompson, CEL, Couceiro, F, Fones, GR et al. In situ flume measurements of resuspension in the North Sea. Estuarine, Coast Shelf Sci 2011;94: 7788.Google Scholar
Ringold, P. Burrowing, root mat density, and the distribution of fiddler crabs in the eastern United States. J Exp Mar Biol Ecol 1979;36: 1121.Google Scholar
Teal, L, Bulling, M, Parker, E, Solan, M. Global patterns of bioturbation intensity and mixed depth of marine soft sediments. Aquat Biol 2010;2: 207218.Google Scholar
Solan, M, Kennedy, R. Observation and quantification of in situ animal-sediment relations using time-lapse sediment profile imagery (t-SPI). Mar Ecol Prog Ser 2002;228: 179191.Google Scholar
Solan, M, Wigham, B, Hudson, I et al. In situ quantification of bioturbation using time-lapse fluorescent sediment profile imaging (f-SPI), luminophore tracers and model simulation. Mar Ecol Prog Ser 2004;271: 112.Google Scholar
Murray, F, Solan, M, Douglas, A. Effects of algal enrichment and salinity on sediment particle reworking activity and associated nutrient generation mediated by the intertidal polychaete Hediste diversicolor. J Exp Mar Biol Ecol 2017;495: 7582.Google Scholar
Germano, J, Rhoads, D, Valente, R et al. The use of sediment profile imaging (SPI) for environmental impact assessments and monitoring studies: Lessons learned from the past four decades. Ocn. and Mar. Biol.: An Annual Rev. 2011;49: 235298.Google Scholar
Schiffers, K, Teal, LR, Travis, JMJ Solan, M. An open source simulation model for soil and sediment bioturbation. PloS one 2011;6: e28028.Google Scholar
Wood, C., Hawkins, S., Godbold, J. et al. Coastal Biodiversity and Ecosystem Service Sustainability (CBESS) Macrofaunal Community Metrics - Total Abundance (TA), Total Biomass (TB), Species Richness (SR), Evenness (J) and Community Bioturbation Potential (BPc) in Mudflat and Saltmarsh Habitats. NERC Environmental Information Data Centre, 2015.Google Scholar
Bishop, C, Skafel, M, Nairn, R. Cohesive profile erosion by waves. Coastal Engineering 1992. American Society of Civil Engineers, 1993, https://doi.org/10.1061/9780872629332.227.Google Scholar
Skafel, MG. Laboratory measurement of nearshore velocities and erosion of cohesive sediment (till) shorelines. Coast Eng 1995;24: 343349.Google Scholar
Skafel, MG, Bishop, CT. Flume experiments on the erosion of till shores by waves. Coast Eng 1994;23: 329348.Google Scholar
Allen, J. R. L.. Mixing at turbidity current heads, and its geological implications: ERRATUM. SEPM J Sediment Res 1971;41: 889.Google Scholar
Thompson, CEL, Amos, CL. The impact of mobile disarticulated shells of Cerastoderma edulis on the abrasion of a cohesive substrate. Estuaries 2002;25: 204214.Google Scholar
Amos, CL, Daborn, G., Christian, H. et al. In situ erosion measurements on fine-grained sediments from the Bay of Fundy. Mar Geol 1992;108: 175196.Google Scholar
Bagnold, RA. The Physics of Blown Sand and Desert Dunes. Methuen, New York.Google Scholar
Amos, C., Sutherland, T, Cloutier, D, Patterson, S. Corrasion of a remoulded cohesive bed by saltating littorinid shells. Cont Shelf Res 2000;20: 12911315.Google Scholar
Thompson, CEL, Amos, CL. Effect of sand movement on a cohesive substrate. J Hydraul Eng 2004;130: 11231125.Google Scholar
Quaresma, V da S, Amos, CL, Bastos, AC. The influence of articulated and disarticulated cockle shells on the erosion of a cohesive bed. J Coast Res 2007;236: 14431451.Google Scholar
Ford, H, Garbutt, A, Ladd, C et al. Soil stabilization linked to plant diversity and environmental context in coastal wetlands. J Veg Sci: Off Organ Int Assoc Veg Sci 2016;27: 259268.Google Scholar
Chen, XD, Zhang, CK, Paterson, DM et al. Hindered erosion: The biological mediation of noncohesive sediment behavior: eps mediating sediment erosion. Water Resour Res 2017;53: 47874801.Google Scholar
Huettel, M, Rusch, A. Advective particle transport into permeable sediments-evidence from experiments in an intertidal sandflat. Limnol Oceanogr 2000;45: 525533.Google Scholar
Howes, BL, Howarth, RW, Teal, JM Valiela, I. Oxidation-reduction potentials in a salt marsh: Spatial patterns and interactions with primary production1: Salt marsh redox potentials. Limnol Oceanogr 1981;26: 350360.Google Scholar
Kristensen, E. Organic matter diagenesis at the oxic/anoxic interface in coastal marine sediments, with emphasis on the role of burrowing animals. Hydrobiologia 2000; 426: 124.Google Scholar
Gerwing, TG, Gerwing, AM, Hamilton, DJ et al. Apparent redox potential discontinuity (aRPD) depth as a relative measure of sediment oxygen content and habitat quality. Int J Sediment Res 2015;30: 7480.Google Scholar
Fanjul, E, Escapa, M, Montemayor, D et al. Effect of crab bioturbation on organic matter processing in South West Atlantic intertidal sediments. J Sea Res 2015;95: 206216.Google Scholar
Kostka, JE, Gribsholt, B, Petrie, E et al. The rates and pathways of carbon oxidation in bioturbated saltmarsh sediments. Limnol Oceanogr 2002;47: 230240.Google Scholar
Negrin, VL, Spetter, CV, Asteasuain, RO et al. Influence of flooding and vegetation on carbon, nitrogen, and phosphorus dynamics in the pore water of a Spartina alterniflora salt marsh. J Environ Sci 2011;23: 212221.Google Scholar
Patrick, WH, Delaune, RD. Nitrogen and phosphorus utilization by Spartina alterniflora in a salt marsh in Barataria Bay, Louisiana. Estuar Coast Mar Sci 1976;4: 5964.Google Scholar
Gutiérrez, JL, Jones, CG, Groffman, PM et al. The contribution of crab burrow excavation to carbon availability in surficial salt-marsh sediments. Ecosystems 2006;9: 647658.Google Scholar
Kristensen, E, Kostka, JE. Macrofaunal burrows and irrigation in marine sediment: Microbiological and biogeochemical interactions. Interactions between Macro‐ and Microorganisms in Marine Sediments. American Geophysical Union, 0, 125–157.Google Scholar
Luckenbach, MW. Sediment stability around animal tubes: The roles of hydrodynamic processes and biotic activity1: Stability around tubes. Limnol Oceanogr 1986;31: 779787.Google Scholar
Talley, TS, Crooks, JA, Levin, LA. Habitat utilization and alteration by the invasive burrowing isopod, Sphaeroma quoyanum, in California salt marshes. Mar Biol 2001;138: 561573.Google Scholar
Hale, R, Mavrogordato, MN, Tolhurst, TJ et al. Characterizations of how species mediate ecosystem properties require more comprehensive functional effect descriptors. Sci Reports 2014;4: 6463.Google Scholar
Hale, R, Boardman, R, Mavrogordato, MN et al. High-resolution computed tomography reconstructions of invertebrate burrow systems. Sci Data 2015;2: 150052.Google Scholar
Coelho, V, Cooper, R, de Almeida Rodrigues, S. Burrow morphology and behavior of the mud shrimp Upogebia omissa (Decapoda: Thalassinidea: Upogebiidae). Mar Ecol Prog Ser 2000;200: 229240.Google Scholar
Yunusa, IAM, Braun, M, Lawrie, R. Amendment of soil with coal fly ash modified the burrowing habits of two earthworm species. Appl Soil Ecol 2009;42: 6368.Google Scholar
Salvo, F, Dufour, SC, Archambault, P et al. Spatial distribution of Alitta virens burrows in intertidal sediments studied by axial tomodensitometry. J Mar Biol Assoc United Kingd 2013;93: 15431552.Google Scholar
Rhoads, DC, Cande, S. Sediment profile camera for in situstudy of organism-sediment relations. Limnol Oceanogr 1971;16: 110114.Google Scholar
Rhoads, D, Germano, J. Characterization of organism-sediment relations using sediment profile imaging: An efficient method of remote ecological monitoring of the seafloor (Remots™ System). Mar Ecol Prog Ser 1982;8: 115128.Google Scholar
Nickell, L, Atkinson, R. Functional morphology of burrows and trophic modes of three thalassinidean shrimp species, and a new approach to the classification of thalassinidean burrow morphology. Mar Ecol Prog Ser 1995;128: 181197.Google Scholar
Koo, BJ, Kwon, KK, Hyun, J-H. The sediment-water interface increment due to the complex burrows of macrofauna in a tidal flat. Ocean Sci J 2005;40, DOI: 10.1007/BF03023522.Google Scholar
Seike, K, Goto, R. Combining in situ burrow casting and computed tomography scanning reveals burrow morphology and symbiotic associations in a burrow. Mar Biol 2017;164, DOI: 10.1007/s00227-017-3096-y.Google Scholar
Oug, E, Høisœter, T. Soft-bottom macrofauna in the high-latitude ecosystem of Balsfjord, northern Norway: Species composition, community structure and temporal variability. Sarsia 2000;85: 113.Google Scholar
Downie, H, Holden, N, Otten, W et al. Transparent soil for imaging the rhizosphere. PloS one 2012;7: 44276.Google Scholar
Rosenberg, R, Davey, E, Gunnarsson, J et al. Application of computer-aided tomography to visualize and quantify biogenic structures in marine sediments. Mar Ecol Prog Ser 2007;331: 2334.Google Scholar
Rosenberg, R, Grémare, A, Duchêne, J et al. 3D visualization and quantification of marine benthic biogenic structures and particle transport utilizing computer-aided tomography. Mar Ecol Prog Ser 2008;363: 171182.Google Scholar
Delefosse, M, Kristensen, E, Crunelle, D et al. Seeing the unseen--bioturbation in 4D: Tracing bioirrigation in marine sediment using positron emission tomography and computed tomography. PloS One 2015;10: e0122201.Google Scholar
Perez, KT, Davey, EW, Moore, RH et al. Application of computer-aided tomography (CT) to the study of estuarine benthic communities. Ecol Appl 1999;9: 10501058.Google Scholar
Mermillod-Blondin, F, Marie, S, Desrosiers, G et al. Assessment of the spatial variability of intertidal benthic communities by axial tomodensitometry: Importance of fine-scale heterogeneity. J Exp Mar Biol Ecol 2003;287: 193208.Google Scholar
Michaud, E, Desrosiers, G, Long, B et al. Use of axial tomography to follow temporal changes of benthic communities in an unstable sedimentary environment (Baie des Ha! Ha!, Saguenay Fjord). J Exp Mar Biol Ecol 2003;285–286: 265282.Google Scholar
Dufour, SC, Desrosiers, G, Long, B et al. A new method for three-dimensional visualization and quantification of biogenic structures in aquatic sediments using axial tomodensitometry: CT scan of biogenic structures. Limnol Oceanogr Methods 2005;3: 372380.Google Scholar
Feagin, RA, Lozada-Bernard, SM, Ravens, TM et al. Does vegetation prevent wave erosion of salt marsh edges? Proc Natl Acad Sci United States Am 2009;106: 1010910113.Google Scholar
Tengbeh, GT. The effect of grass roots on shear strength variations with moisture content. Soil Technol 1993;6: 287295.Google Scholar
Mamo, M., Bubenzer, G. D.. Detachment rate, soil erodibility, and soil strength as influenced by living plant roots part i: Laboratory study. Trans ASAE 2001;44: 11671174.Google Scholar
Chen, Y, Thompson, CEL, Collins, MB. Saltmarsh creek bank stability: Biostabilisation and consolidation with depth. Cont Shelf Res 2012;35: 6474.Google Scholar
De Baets, S, Poesen, J, Gyssels, G Knapen, A. Effects of grass roots on the erodibility of topsoils during concentrated flow. Geomorphology 2006;76: 5467.Google Scholar
Khanal, A, Fox, GA. Detachment characteristics of root-permeated soils from laboratory jet erosion tests. Ecol Eng 2017;100: 335343.Google Scholar
Katuwal, S, Vermang, J, Cornelis, WM et al. Effect of root density on erosion and erodibility of a loamy soil under simulated rain. Soil Sci 2013;178: 2936.Google Scholar
Ghidey, F., Alberts, E. E.. Plant root effects on soil erodibility, splash detachment, soil strength, and aggregate stability. Trans ASAE 1997;40: 129135.Google Scholar
Painting, SJ, van der Molen, J, Parker, ER et al. Development of indicators of ecosystem functioning in a temperate shelf sea: A combined fieldwork and modelling approach. Biogeochemistry 2013;113: 237257.Google Scholar
Queirós, AM, Birchenough, SNR, Bremner, J et al. A bioturbation classification of European marine infaunal invertebrates. Ecol Evol 2013;3: 39583985.Google Scholar
Steiner, N, Deal, C, Lannuzel, D et al. What sea-ice biogeochemical modellers need from observers. Elem Sci Anthr 2016;4, p.000084. DOI: 10.12952/journal.elementa.000084.Google Scholar
Morrisey, D, Howitt, L, Underwood, A et al. Spatial variation in soft-sediment benthos. Mar Ecol Prog Ser 1992;81: 197204.Google Scholar
Savchuk, OP. Nutrient biogeochemical cycles in the Gulf of Riga: Scaling up field studies with a mathematical model. J Mar Syst 2002;32: 253280.Google Scholar
Thompson, CEL, Silburn, B, Williams, ME et al. An approach for the identification of exemplar sites for scaling up targeted field observations of benthic biogeochemistry in heterogeneous environments. Biogeochemistry 2017;135: 134.Google Scholar
Kirwan, ML, Mudd, SM. Response of salt-marsh carbon accumulation to climate change. Nature 2012;489: 550553.Google Scholar
Morris, JT, Sundareshwar, PV, Nietch, CT et al. Responses of coastal wetlands to rising sea level. Ecology 2002;83: 28692877.Google Scholar
Langley, JA, McKee, KL, Cahoon, DR et al. Elevated CO2 stimulates marsh elevation gain, counterbalancing sea-level rise. Proc Natl Acad Sci United States Am 2009;106: 61826186.Google Scholar
Cherry, JA, McKee, KL, Grace, JB. Elevated CO2 enhances biological contributions to elevation change in coastal wetlands by offsetting stressors associated with sea-level rise. J Ecol 2009;97: 6777.Google Scholar
Kirwan, ML, Guntenspergen, GR, Morris, JT. Latitudinal trends in Spartina alterniflora productivity and the response of coastal marshes to global change. Glob Chang Biol 2009;15: 19821989.Google Scholar
Charles, H, Dukes, JS. Effects of warming and altered precipitation on plant and nutrient dynamics of a New England salt marsh. Ecol Appl: Publ Ecol Soc Am 2009;19: 17581773.Google Scholar
Mudd, SM, Howell, SM, Morris, JT. Impact of dynamic feedbacks between sedimentation, sea-level rise, and biomass production on near-surface marsh stratigraphy and carbon accumulation. Estuarine, Coast Shelf Sci 2009;82: 377389.Google Scholar
Kirwan, ML, Guntenspergen, GR. Feedbacks between inundation, root production, and shoot growth in a rapidly submerging brackish marsh: Marsh root growth under sea level rise. J Ecol 2012;100: 764770.Google Scholar
Duarte, CM, Losada, IJ, Hendriks, IE et al. The role of coastal plant communities for climate change mitigation and adaptation. Nat Clim Chang 2013;3: 961968.Google Scholar
Donnelly, JP, Bertness, MD. Rapid shoreward encroachment of salt marsh cordgrass in response to accelerated sea-level rise. Proc Natl Acad Sci United States Am 2001;98: 1421814223.Google Scholar
Hartig, EK, Gornitz, V, Kolker, A et al. Anthropogenic and climate-change impacts on salt marshes of Jamaica Bay, New York City. Wetlands 2002;22: 7189.Google Scholar
Scavia, D, Field, JC, Boesch, DF et al. Climate change impacts on U.S. Coastal and marine ecosystems. Estuaries 2002;25: 149164.Google Scholar
Farron, SJ, Hughes, Z, FitzGerald, DM, and Storm, KB. The impacts of bioturbation by common marsh crabs on sediment erodibility: A laboratory flume investigation. Estuarine, Coastal and Shelf Science 2020;238: 111.Google Scholar
Asano, T, Tsutsui, S, and Sakai, T. 1988. Wave damping characteristics due to seaweed. Proceedings of the 35th Coastal Engineering Conference in Japan. JSCE. 138–142.Google Scholar

Save book to Kindle

To save this book to your Kindle, first ensure coreplatform@cambridge.org is added to your Approved Personal Document E-mail List under your Personal Document Settings on the Manage Your Content and Devices page of your Amazon account. Then enter the ‘name’ part of your Kindle email address below. Find out more about saving to your Kindle.

Note you can select to save to either the @free.kindle.com or @kindle.com variations. ‘@free.kindle.com’ emails are free but can only be saved to your device when it is connected to wi-fi. ‘@kindle.com’ emails can be delivered even when you are not connected to wi-fi, but note that service fees apply.

Find out more about the Kindle Personal Document Service.

Available formats
×

Save book to Dropbox

To save content items to your account, please confirm that you agree to abide by our usage policies. If this is the first time you use this feature, you will be asked to authorise Cambridge Core to connect with your account. Find out more about saving content to Dropbox.

Available formats
×

Save book to Google Drive

To save content items to your account, please confirm that you agree to abide by our usage policies. If this is the first time you use this feature, you will be asked to authorise Cambridge Core to connect with your account. Find out more about saving content to Google Drive.

Available formats
×